Skip to main content

Full text of "Modern Science And Its Philosophy"

See other formats



DELHI UNIVERSITY LIBRARY 

CK No. X;' "X • /? fl // 

— - ^ " Date of release for loan 

Ac. No. f 7 ^ p ' - ' ■ 

'Vbis book should be returned on or before the date last Stamped 
beluw. An overdue charge of Six nP. will be charged for each dajr 
the book is kept overtime. 




modem science 


and its philosophy 





PHILtPP FRANK 


modern science 
and its philosophy 


HARVARD UNIVERSITY PRESS • CAMBRIDGE • 1949 

LONDON -GEOFFREY CUMBERIEGE • OXFORD UNIVERSITY PRESS 




COPYRIGHT 
1941 • 1949 

BY THE PRESIDENT AND PELIOWS OF HARVARD COLLEl 
PRINTED IN THE UNITED STATES OF AMERICA 



DEDICATED TO 

MANIA 


IN REMEMBRANCE. OF THE FAST 
AND ANTICIPATION 
OF THE FUTURE 




PREFACE 


Along with the evolution of twentieth-century science, philosophi- 
cal ideas were developed which have sprouted and grown on the soil 
of this science. In this book I present my thoughts on this develop- 
ment, in the sequence in which they entered my mind during the 
many years of my teaching and research work in science. If we want 
to evaluate precisely and critically how firmly this philosophy is an- 
chored in the ground of science, we must not ignore the extrascientific 
factors, but must analyze carefully the social, ethical and religious in- 
fluences. Every satisfactory philosophy of science has to combine logic 
of science with sociology of science. 

Since 1940 not only have I been teaching science, but Harvard 
University offered me the opportunity of teaching also the philosophy 
which, in my opinion, has grown up along with twentieth-century 
science. A large part of this book reflects the experience that I have 
% gained by observing the interests of the students and their reactions 
to my teaching. 

My work has been stimulated greatly by the Harvard experiment 
in General Education, in which I have had the privilege of participat- 
ing. The Harvard University Press has been of substantial help to me 
in finishing this book. I am particularly obliged to Miss Eleanor R. 
Dobson and Mr. J. D. Elder for their fine spirit of cooperation. I hope 
that the Harvard Press will soon publish a volume which is written 
along similar lines, the lucid and straightforward book ‘‘Positivism,” 
by R. von Mises. 


PmLTPP Frame 




CONTENTS 


^ introduction . historical background - J 

<1. experience and the law of causality 53 

j2.' the importance for our times of Ernst Mach's philosophy of 

science 61 

3. Ernst Mach and the unity of science 79 

4. physical theories of the twentieth century and school philosophy 90 

5. is there a trend today toward idealism in physics? 122 

6. mechanical "explanation" or mathematical description? 138 

7. modern physics and common sense 144 

8. philosophic misinterpretations of the quantum .theory 158 

9. determinism and indeterminism in modern physics 172 

10. how idealists and materialists view modern physics 186 

11. logical empiricism and the philosophy of the Soviet Union 198 

12. why do scientists and philosophers so often disagree about the 

merits of a new theory? 207 

•13. the philosophic meaning of the Copernican revolution 216 

14. the place of the philosophy of science in the curriculum of the 

physics student 228 

science teaching and the humanities 260 

/1 6. the place of logic and metaphysics in the advancement of modern 

science 286 

bibliographical note 305 

index 307 




INTRODUCTION 


HISTORICAL BACKGROUND 


1. Discusilons in a Vienna CofFee House 

At the time when the first chapter of this book was written (190f ) 
I had just graduated from the University of Vienna as a doctor of phi- 
losophy in physics. But the domain of my most intensive interest was 
the philosophy of science. 1 used to associate with a group of students 
who assembled every Thursday night in one of the old Viennese coffee 
houses. We stayed until midnight and even later, discussing prob- 
lems of science and philosophy. Our interest was spread widely over 
many fields, but we returned again and again to our central prob- 
lem: How can we avoid the traditional ambiguity and obscurity of 
philosophy? How can we bring about the closest possible rapproche- 
ment between philosophy and science? By “science” we did not mean 
“natural science” only, but we included always social studies and the 
humanities. The most active and regular members of our group were, 
besides myself, the mathematician, Hans Hahn, and the economist, 
Otto Neurath. 

Although all three of us were at that time actively engaged in re- 
search in our special fields, we made great efforts to absorb as much 
information, methodology and background from other fields as we 
were able to get. Our field of interest included also a great variety of 
political, historical, and religious problems which we discussed as 
scientifically as possible. Our group had at that time no particular 
common predilection for a certain political or religious creed. We had, 
however, an inclination towards empiricism on one hand and long 
and clear-cut chains of logical conclusions on the other. There were 
quite a few occasions on which these two predilections did not mix 
very well. 

This apparent internal discrepancy provided us, however, with a cer- 
tain breadA of approach by which we were able to have helpful discus- 



modern science and its philosophy 


sions with followers of various philosophical opinions. Among the par- ‘ 
ticipants in our discussions were, for instance, several advocates of 
Catholic philosophy. Some of them were Thomists, some were rather 
adherents of a romantic mysticism. Discussions about the Old and New 
Testaments, the Jewish Talmud, St. Augustine, and the medieval 
schoolmen were frequent in our group. Otto Neurath even enrolled for 
one year in the Divinity School of the University in order to get an ade- 
quate picture of Catholic philosophy, and won an award for the best 
paper on moral theology. This shows the high degree of our interest 
in the cultural background of philosophic theories and our belief in 
the necessity of an open mind which would enable us to discuss our 
problems with people of divergent opinions. 

At that time a prominent French historian and philosopher of sci- 
ence, Abel Rey, published a book which later was to make a great im- 
pression upon me. At the turn of the century the decline of mechanistic 
physics was accompanied by a belief that the scientific method itseli^ 
had failed to give us the “truth about the universe”; hence nonscientific 
and even antiscientific tendencies gained momentum. I quote some 
passages in which Rey describes this situation excellently and pre- 
cisely. 

Fifty years ago, he says, the explanation of nature was believed to 
be purely mechanical. 

It was postulated that physics was nothing but a complication of me- 
chanics, a molecular mechanics . . . Today [1907] it seems that the picture 
offered by the physical sciences has changed completely. The general unity 
is replaced by an extreme diversity, not only in the details, but in the leading 
and fundamental ideas . . . [This accounts for] what is called the crisis 
of contemporary physics. Traditional physics assumed until the middle of 
the nineteenth centiny that it had only to continue its own path to become 
the metaphysics of matter. It ascribed to its theories an ontologic value, and 
these theories were all mechanistic. Traditional mechanistic physics was 
supposed, above and beyond the results of experience, to be the red, cog- 
nition of the material universe. This conception was not a hypothetical 
description of our experience; it was a dogma. 

The criticism of the traditional mechanistic physics that was formu- 
lated in the second half of the nineteenth century weakened this assertion 
of the ontologic reality of mechanistic physics. Upon this criticism a philos- 


2 



historical background 


*ophy of^hysics was established that became almost traditional toward the 
end of the nineteenth century. Science became nothing but a symbolic pat- 
tern, a frame of reference. Moreover, since this frame of reference varied 
according to the school of thought, it was soon discovered that actually noth- 
ing was referred that had not previously been fashioned in such a way that 
it could be so referred. Science became a work of art to the lover of pure 
science, a product of artisanship to the utilitarian. This attitude could quite 
rightly be interpreted as denying the possibility that science can exist. A 
science that has become simply a useful technique ... no longer has the 
right to call itself science without distorting the meaning of the word. To 
say that science cannot be anything but this means to negate science in the 
proper sense of the word. 

The failure of the traditional mechanistic science . . . entails the prop- 
osition: “Science itself has failed.” . . . We can have a collection of em- 
pirical recipes, we can even systematize them for the convenience of 
memorizing them, but we have no cognition of the phenomena to which this 
system or these recipes are applied.* 

Our group was formed during the period which was so eloquently 
described by Rey. His book was discussed frequently by us in the last 
years of my stay in Vienna (1908-1912). The problems raised and the 
results obtained are reflected partly in Chapter 2 of this book. The 
general reaction of our group to the intellectual and cultural situation 
depicted by Rey can be described as follows: 

We recognized the gradual decline in the belief that mechanistic sci- 
ence would eventually embrace all our observations. This belief had 
been closely connected with the belief in progress in science and in 
the scientific conception of the world. Therefore, this decline brought 
about a noticeable uneasiness. Many people lost their faith in scientific 
method and looked for some other method which might yield a real 
understanding of the world. A great many people believed, or at least 
wanted to believe, that the time had come to return to the medieval 
ideas that may be characterized as the organismic conception of the 
world. 

In the history of science and philosophy there have always been di- 
vergent opinions about the conditions under which we may say that a 

*A. Rey, La ThSorie de physique chez les physiclens contemporains (Paris, 
1907), pp. 16 £F. 


3 



modern science and its philosophy 


scientific theory has “explained” a certain range of observations. Some^ 
authors have maintained that only an explanation by mechanical causes 
and by the motion of material particles can satisfy our intellectual curi- 
osity. Others have claimed that the reduction to mechanical causes is 
only a superficial explanation and not a real one. Some of the opponents 
of the mechanistic world view have stated that all phenomena must 
b,e interpreted in terms of the evolution of an “organic whole” in order 
really to understand them. The decline of the belief in mechanistic sci- 
ence seemed to favor this organismic view, which has been attractive 
to many because of its religious and social implications. In this way 
there had arisen at the turn of the century what some called a crisis in 
science or, more accurately, in the scientific conception of the world. 
For more than two centuries the idea of progress in science and human 
life had been connected with the advance of the mechanistic explana- 
tion of natural phenomena. Now science itself seemed to abandon this 
mechanistic conception, and the paradoxical situation arose that one 
could fight the scientific conception of the world in the name of the 
advance of science. 


2« The Failure of Mechanistic Science 

The sixteen chapters of this book have been written over a period of 
almost forty years. They are all meant to be contributions to one task: 
to break through the wall which has separated science and pliilosophy 
for about one and one-half centuries. The book reflects the methods by 
which this wall has been besieged in the twentieth century. During the 
last decades of the nineteenth century, a revolution in physical science 
started which has itself brought about a revolution in our general scien- 
tific thought. The methods that have been tried in the fight for the unity 
between science and philosophy have changed along with the advance 
of science.^ wo characteristic beliefs of nineteenth-century science 
broke down during its last decades; these were the belief that all phe- 
nomena in nature can be reduced to th e laws of mechani cs, and the 
belief that science will eventually reveal the “truth” about the univers^ 
In the twentieth century the revolutionary changes in science de- 
veloped with ever-increasing rapidity and intensity. It is no wonder 


4 



historical background 


*that thijf rapid transformation of scientific thought has been accom- 
panied by rapidly changing methods in the scientist’s approach to phi- 
losophy and in the philosopher’s views on science. 

The chapters of this book have played their part in this developing 
and changing fight for the unity of science and philosophy. Through 
them we can pursue a pattern that has grown from rather tentative 
and naively empirical beginnings to a more and more abstract tech- 
nique. In order to understand this evolution precisely, it is not su:^- 
cient to follow the gradual alterations from the purely logical angle. 
We must also consider carefully the historical trend and background 
of the arguments. 

Rey had the strong feeling that the place of physics in human 
thought had dangerously deteriorated. He says: 

The physicochemical sciences are an effort made by man to explain 
sensed nature or what is perceived by our senses . . . The importance of 
this effort has been understood in all periods of history. So early a thinker 
as Epicurus believed that physics was the basis of the effort to liberate 
the human mind from its blind instinct to believe, from its prejudices and 
its superstitions . . . Everyone would agree that^modem positivism has 
been nothing but an attempt to extend to all the departments of human 
knowledge, without exception, the general method and conception of sci- 
ence, that is, science constructed according to the example of physicsj)the 
spirit of positivism and the spirit of science have become synonymous in 
current speech. 

If these sciences, which historically have played essentially an emancipa- 
tory role, were tarnished by a crisis that leaves them no other value than that 
of recipes which are technically useful and that deprives them of any sig- 
nificance in the cognition of nature, a complete upset in the art of logic and 
in the history of ideas must result.(^hysics loses all its educational value; 
the positive spirit which it represent is a misleading and dangerous spirit. 
Reason, rational method, and experimental method must be considered in 
good faith as having no cognitive value. All these methods are, then, pro- 
cedures of action, not means of cognition. They can be developed in order 
to obtain certain practical results, but we must be well aware that they have 
no value except in their restricted domain. The cognition of the real must 
be sought or be given by other means. We must guard against the danger- 
ous illusion of rationalism and scientism. It is important to know that by this 
method the real is ignored and that physics leads to ignorance and not to 
cognition of real nature . . ]) 


5 



modern science and Its philosophy 


If the problem of the cognition of nature and of the posslbili^ of ther 
physicochemical sciences remains in the same form in which it has devel- 
oped from the Renaissance to the time of positivism, the rational and posi- 
tive method remains the supreme educator of the human mind, in the 
domain that is accessible to it, of course, T o give the mind a scientific atti- 
tude in the sense that has been understood by positivism and positive 
physics remains the necessary and sufficient condition of intellectual sanity. 
Physics is the school where one learns to know things.^ 

( 

3. Ernst Mach's Purge of Traditional Physics 

,jjn this critical situation our minds turned towards a solution that 
had been advanced about twenty-five years before by our local 
physicist and philosopher, Ernst Mach. He maintained that “explana- 
tion" by reduction to a system of cherished conceptions is pure illu- 
sion. If all the multitude of observable phenomena are reduced to 
mechanical or organismic phenomena, these special types of phe- 
nomena chosen as the basis of explanation are by themselves no more 
understandable than the phenomena that are to be explained. Maeh 
claimed that there is no essential difference be tween a n “e xplanation" 

1 and a "desc ription." In our everyday language the word “description” 

I refers to a single ev ent or phenomenon. We^ “describe” the fall of a 
specific stone at a specific time and place. If, however, we formulate 
Galileo’s law of freely falling bodies, which tells us that all bodies fall 
with an equal constant acceleration, we describe the fall of a great 
-many bodies under different circumstances. This description of a_ 
class of phenomena is called by Mach a “physical law” or an “explana- 
tion.” If an explanation is nothing but a^ descripti on of many cases 
1 by a short sentence, it cannot matter much whether the "explanation” 
is given by feduCfioh to the laws of mechanics or to the laws of electro- 
I magnetism or even of statistics.^ 

' In this way Mach separated the conception of “scientific explana- 
tion” from “mechanical explanation.” He saved the scientific world pic- 
ture from going down along with the mechanistic picture. Similar ideas 
had been advanced by other writers in Mach’s time, but none formu- 
lated them with so much lucidity and breadth of approach. No one else 

* Ibid., pp. 18 ff. 



historical background 


* ancho];^ d tliem so firmly in the soil of science, in physics as well as in 
biology and psychology. 

Although Mach’s views were the principal background of our 
thoughts, they were not the most powerful stimulus to our actual 
work. Our group fully approved Mach’s antimetaphysical tendencies, 
and we joined gladly in his radical empiricism as a starting point; 
(but we felt very strongly about the primary role of mathematics and 
logic in the structure of science. It seemed to us that Mach had not 
done full justice to this aspect of science. We felt that considering the 
principles of science as nothing but abbreviated descriptions of sense 
observations did not account fully for the fact that the principles of 
science contain simple clear-cut mathematical relations among a small 
ni’.mber of concepts, whereas every description of observations con- 
tains a great number of vague connections among a great number of 
vague concepts. We also felt that to call the principles of science “eco- 
nomic descriptions of observations” was not to do justice to the pre- 
dominant role of reasoning in the discovery and presentation of these 
principles^ We were even attracted by some tenets of Kant’s theory of 
knowledge, particularly by his “Introduction to Any Future Meta- 
physics that may Present Itself as Science.” ^e saw much truth in 
Kant’s statement that the recording of observations is not a purely 
passive act but that a great deal of mental activity is necessary in order 
to formulate general statements about sense observations.) ' 

The human mind, according to KanL-can desc ribe n atural phenom- 
ena only by using a certain j)a ttero, c ertain forms of thinking that a re 
produced by the observing mmd an3~are~not prwided by the ob- 
served physical object. Since these “forms" or “patterns of experience” 
are provided by the human mind and not by the physical facts, they 
cannot be changed by the advance of scientific investigations^The ma- 
terial delivered by sense observation can be manufactured into a scien- 
tific system only by adding forms of experience that are determined by 
the nature of the human mind^ur whole group understood and fully 
agreed that the human mind is ^rtly responsible for the content of sci- 
entific propositions and theories. But we also felt strongly that an 
orthodox belief in Kant would lead us into a new type of metaphysics. 


7 



modern science and its philosophy 


not less unscientific than medieval metaphysics. For instancy, Kant, ' 
and even more, many of his followers, maintained that the axioms of 
Euclidean geometry or Newton’s laws of motion are forms of experi- 
ence that are produced by the organization of the human mind. This 
would imply that no advance in these fundamentals of science will be 
possible without changing the nature of human mind. However, about 
1900, scientists became accustomed to envisaging the possibility of! 
changes in the principles that had been traditionally regarded as self- 
evident axioms^ Non-Euclidean geometry was no longer considered 
a purely logical exercise, but was recognized as a serious system of 
geometry. This new conception, along with Mach’s criticism of New- 
ton’s mechanics and with the new electromagnetic theory of matter, 
opened a path for some(^doubts whether Newton’s laws of motion were 
actually the final 'word about the nature of the universe. A step in the 
same direction was the statistical interpretation of thermodynamics, 
which suggested the superposition of statistical laws on dynamical 
laws. As enthusiastic students of contemporary science our group re- 
jected Kant’s doctrine that the forms of experience provided by the 
human mind were unchangeable. We looked for some way to construe 
these forms as subject to an evolution that would be in accord with the 
evolution of science. We felt very strongly that there was a certain gap 
between the description of observations, necessarily vague and com- 
plex, and the principles of science, consisting, in physics particularly, 
of a small number of concepts (like force, mass, etc.) linked by state- 
ments of great simplicity.^We admitted that the gap between the 
description of facts and the general principles of science was not fully 
bridged by Mach, but we could not agree with Kant, who built this 
bridge by forms or patterns of experience that could not change with 
the advance of scienc^ 


4. Henri Poincare and a ''New Positiviim" 

In our opinion, the man who bridged the gap successfully was the 
French mathematician and philosopher Henri Poincar6. For us, he 
was a kind of Kant freed of the remnants of medieval scholasticism 
and anointed with the oil of modem science. 


8 



historical background 


^ Kaat had claimed that there would never be another way to organ- 
ize our experience than by Euclidean geomet ry and Newtonian me- 
chanics. But at the turn of the century non-Euclidean geometry had 
been established, although its importance for physical science was not 
completely understood. Departures from Newtonian mechanics were 
already in the making. We understood that Kant’s erroneous conception 
of geometry and mechanics must have its source in his erroneous ajti- 
tude toward the relation between scifincft and philosophy . ) 

At this time I was very much interested in the criticism launched 
by Nietzsche against Kant’s idealistic philosophy. Kant’s primary aim j 
was to answer the question of how the human mind can make state- 1 
ments about facts of the external world with absolute certainty even if 
these assertions are not the result of experience about this world. For 
Kant, the propositions of Euclidean geometry were a convincing ex- 
ample of assertions of this kind. By claiming that the axioms of geom- 
etry are forms of experience produced by the human mind, Kant ex- 
plained man’s ability to produce these assertions without external ex- 
perience. Nietzsche said flippantly that Kant’s explanation is merely 
equivalent to saying that man can do it “by virtue of a virtue.” Nietzsche . 
accused him of demonstrating by sophisticated and obscure arguments 
that popular prejudices are right while the scientists are wrong. We 
appreciated in Poincare just what was different from Kant. We agreed 
with Abel Bey’s characterization of Poincare’s contribution as a “new 
positivism” which was a definite improvement over the positivism of 
Comte and Mill. Bey wrote: 

The new positivism certainly has its origin in the positivism of Comte, 
of Taine and Mill . . . It is rejuvenated but has preserved the great direc- 
tive ideas from its previous stage: the relativism and empiricism of our 
knowledge. But it stresses . . . the idea of experimental categories which 
are required by science as a central necessity . . . Positivism is renewed by'^ 
building a new rationalism upon the criticism of traditional rationalism inj 
the second half of the nineteenth century . . . What was lacking in Comte’s 
or Mill’s positivism . . . was their . . . failure to have established in a , 
new form a theory of categories. Objective experience is not something which ^ 
is outside and independent of our minds. Objective experience and mind are f' , 
functions of each other, imply each other, and exist by virtue of each other. 


9 



modern science and its philosophy 

‘ c 

To say that the relations between physical objects derive from the nature of 
these objects and to say that these relations are constructed by our minds 
are two artificial theories ... _ 

Our experience is a system, a relation of relations. The relation is the I 
given.®’ 

5. Mach, Poincari and Lanin 

^ Chapter 1 of the present book is written in the spirit of this new 
positivism. Poincare claimed that the general laws of science— the law 
of inertia, the principle of conservation of energy, etc.— are neither state- 
ments about facts which can be checked by experiments, nor a priori 
statements w'hich necessarily emanate from the organization of human 
mind. They are rather arbitrary conventions about how to use some 
words or expressions. In this chapter Poincare’s basic tenets are applied 
to the law of causality. 

I started from Hume’s formulation of this law: When a state A of a 
system is followed one time by a state B, every time that A recurs it is 
followed again by B. I analyzed this formulation in a way that is sim- 
ilar to Poincare’s analysis of the general principles of science. I put the 
question: How do we know when the state A has recurred? There is 
no exact method except to investigate whether it is followed by B, 
Hence the law of causality is not a statement about observable physical 
facts but a definition of the expression, “the state A has recurred.” 

When I first published this paper ( 1907) it aroused a certain amaze- 
ment among scientists. Among the comments were those of two men 
of world- wide fame, although in different fields: Einstein and Lenin. 
Einstein’s letter was my first personal contact with him. He approved 
the logic of my argument, but he objected that it demonstrates only 
that there is a conventional element in the law of causality and not that 
it is merely a convention or definition. He agreed with me that, what- 
ever may happen in nature, one can never prove that a violation of the 
law of causality has taken place. One can always introduce by con- 
vention a terminology by which this law is saved. But it could happen 
that in this way our language and terminology might become highly 
complicated and cumbersome. What is not conventional in the law of 
® Ibid., pp. 392 S. 


10 



historical background 


» y 

‘ causalit;^ is the fact that we can save this law by using a relatively 
simple terminology: we are sure that a state A has recurred when a 
small number of state variables have the same values that they had at 
the start. This “simplicity of nature” is the observable fact which can- 
not be reduced to a convention on how to use some words. These re- 
marks had a great influence on my thought on the future course of the 
philosophy of science. I realized that Poincare’s conventionalism needs 
qualifications. One has to distinguish between what is logically pos- 
sible and what is helpful in empirical science. In other words, logic 
needs a drop of pragmatic oil. 

Lenin’s comment was rather unfavorable. In his book “Materialism 
and Empiriocriticism” (we would call it today “Materialism and Posi- 
tivism”) he maintains that, “as a Kantian,” I “rejoiced” to be able to 
give support to Kant’s idealism by the “most modern philosophy of sci- 
ence.” From my reference to the relations between Poincar^ and Kant . 
he drew the conclusion that I tried to make use of Poincar^ in the serv- 
ice of idealistic philosophy and that, therefore, this paper had an anti- 
materialistic and reactionary tendency. Lenin’s book did not come to 
my attention before the early twenties. Then, however, it stimulated 
me to think over more carefully the relation between Poincar4 and 
Kant, between positivism and ideah'sm and, particularly, to investigate 
the role that metaphysical interpretations of scientific theories have 
played in support of political and religious philosophies. 

However, during the interval ( 1907-1917) between writing the first 
and second chapters of the present book, my interest was directed 
mainly toward any possible advance in the logic of science. I was con- 
vinced that the solution must be sought by starting from the ideas of 
such men as Mach and Poincar6.' 

At first glance these two authors seemed to contradict each other 
flagrantly. I soon realized that any advance in the philosophy of sci- 
ence would consist in setting up a theory in which the views of Mach 
and of Poincare would be two special aspects of one more general view. 

To summarize these two theories in a single sentence, one might say: 
According to Mach the general principles of science are abbreviated 
economical descriptions of observed facts; according to Poincare they 


II 



modern science and its philosophy 

are free creations of the human mind which do not tell anything about* 
observed facts. The attempt to integrate the two concepts into one co- ' 
herent system was the origin of what was later called logical empiri- j 
cism. 


6. D. Hilbert's New Foundations of Geometry 

, The traditional presentation of physical theories frequently consists 
of a system of statements in which descriptions of observations are 
mixed with mathematical considerations in such a way that sometimes 
one cannot distinguish clearly which is which. It is Poincare’s great 
merit to have stressed that one part of every physical theory is a set of 
arbitrary axioms and logical conclusions drawn from these axioms. 
These axioms are relations between signs, which may be words or alge- 
braic symbols; the important point is that the conclusions that we draw 
from these axioms are not dependent upon the meanings of these sym- 
bols. Hence this part of a theory is purely conventional in the sense 
of Poincare. It does not say anything about observable facts, but only 
leads to hypothetical statements of the following type: “If the axioms 
of this system are true, then the following propositions are also true,” 
or still more exactly speaking: “If there is a group of relations between 
these symbols, there are also some other relations between the same 
symbols.” This state of affairs is often described by saying that the 
system of principles and conclusions describes not a content but a 
structure. Hence, this system is occasionally referred to as the struc- 
tural system. 

The simplest example is geometry. It is the first example in the his- 
tory of science of a logical system that claims to make statements about 
facts in the observable world as well. Geometry did not proceed ac- 
cording to the pattern adopted by the older positivists like Hume, 
Comte, or Mill. It did not collect facts and draw conclusions from 
these observations. Instead, it built up a system of axioms which were 
statements containing abstract terms like “point,” “straight line,” “inter- 
section.” Conclusions from these axioms can actually be dravm without 
knowing the meaning of these terms. In the textbooks of geometry that 
were written in accordance with the tradition laid down by Euclid, 


12 



historical background 


'such tevns as “straight line” or “intersection” seemed to have a “mean- 
ing” in the same sense as the words “table” or “horse” have a meaning, 
except that the geometric concepts were supposed to be the names of 
“idealized” physical objects. 

After the turn of the century, however, David Hilbert started a 
purge of the foundations of geometry and set up a clear-cut and con- 
sistent system of axioms. He stressed that such concepts as “point” ^r 
“straight line” have no meaning besides the one defined by the axioms. 
The axioms are “implicit definitions” of the geometric concepts. For 
Hilbert, as for Poincare, the axioms were conventions about the use of 
the geometric terms. In this way Hilbert made a significant contribu- 
tion to the “new positivism.” He restricted himself, however, to the in- 
vestigation of the structural system and did not discuss, as Foincar4 
did, the relation of the geometric axioms to our experience or our sense 
observations. The propositions that are derived from the axioms cannot 
contain any word besides the symbols contained in the axioms. But 
these symbols have no meaning in the physical world. The great asset 
of this method is that conclusions can be drawn without being affected 
by the vagueness of terms that describe our actual observations, like 
“red,” "blue,” “warm,” “sweet.” This method of geometry became the 
method of the mathematical physics of the nineteenth century. Heat, 
electricity, and light were described by systems of principles that con- 
sisted not of observational terms but of abstract symbols. However, the 
symbols occurring in the axioms and propositions, of geometry as well 
as of mathematical physics, can be linked to observable facts in a brief 
and easily understandable way; for example, the straightness of a line 
can be checked approximately by the edge of a rigid table, or a vol- 
ume or a temperature can be measured by simple physical methods. 


7. The New Positivism and P. Duhem's "Thomism'' 

The axiomatic or structural system, including its conclusions, is 
merely an arbitrary convention if the purely logical viewpoint is main- 
tained without going into the physical interpretation. It was clear to 
, Poincar^ that the structiural system is logically arbitrary because it can- 
not be demonstrated by logical means. It is not psychologically arbi- 


13 



modern science and its pniiosopny 


trary, however, because in practice we construct only those syst^s that ' 
can be interpreted in terms of physical facts and that are therefore help- 
ful for the formulation of natural laws. 

If this line of reasoning is followed, we can see, in a perfunctory 
way at least, how Mach’s and Poincare’s ideas about the general prin- 
ciples of science can be integrated. The axiomatic system, the set of 
relations between symbols, is a product of our free imagination; it is 
arbitrary. But if the concepts occurring in it are interpreted or identi- 
fied with some observational conceptions, our axiomatic system, if well 
chosen, becomes an economical description of observational facts. 

Now the presentation of the law of causality as an arbitrary con- 
vention (Chapter 1) can be freed of its paradoxical appearance. The 
law of causality, as a part of an axiomatic system, is an arbitrary con- 
vention about the use of terms like “the recurrence of a state of a sys- 
tem,” but if interpreted physically it becomes a statement about ob- 
servable facts. In this way, the philosophy of Mach could be integrated 
into the "new positivism” of men like Hemd Poincare, Abel Hey, and 
Pierre Duhem. The connection between the new positivism and the 
old teaching of Hume and Comte is the requirement that all abstract 
terms of science— such as force, energy, mass— must be interpreted in 
terms of sense observations. Exactly speaking, every statement in 
which these abstract terms occur must be interpreted as a statement 
about observational facts. Mach formulated this requirement by say- 
ing that all scientific statements are statements about sense ob- 
servations. 

Mach’s requirements have frequently been misinterpreted. Some 
authors considered them a kind of subjective idealism, meaning that 
the world consists of sense data only. Others considered them a kind 
of skepticism or agnosticism, meaning that man cannot know anything 
about the true or real world; man and man-made science can tell us 
only about our sense observations, while the objective reality will be 
eternally unknown to human intelligence. In this interpretation science 
would be purely subjective, a collection of statements about subjective 
sense impressions. This misinterpretation had its source in traditional 
philosophy, according to which science and philosophy must find the 


14 



historical background 


idden ti;pasure of truth behind appearances. All the various systems 
■ traditional philosophy seeking objective reality behind appearances 
ere opposed to Mach’s philosophy. Moreover, a great many scientists 
Lsliked the idea that the statements of science do not describe the 
sal world but are only statements about our sense observations, with- 
it objective significance. 

The property of the structural system of not telling us anything 
lOut the world of observable physical facts was particularly empha- 
zed by the French scientist, philosopher, and historian, Pierre Duhem. 
is writings exerted a strong infiuence upon our group and, particu- 
rly, upon my own thinking. By studying Duhem thoroughly we 
lined a more subtle understanding of the relations among science, 
etaphysics, and religion than has been customary among empiricists, 
uhem says, much as Mach had done, 

A theory of physics is not an explanation; it is a system of mathematical 
opositions deduced from a small number of principles the aim of which is 
represent as simply, as completely, and as exactly as possible, a group of 
perimenlal laws.* 

This formulation is a great step on the way toward an integration 
Mach and Poincar4. Duhem understood very well that no single 
•oposition of a physical theory can be said to be verified by a specific 
:periment. The theory as a whole is verified by the whole body of ex- 
jrimental facts. As Duhem put it. 

The experimentum crucis is impossible in physics.' 
e says again; 

The watchmaker to whom one gives a watch that does not run will take it 
[ apart and will examine each of the pieces until he finds out which one 
damaged. The physician to whom one presents a patient cannot dissect 
m to establish the diagnosis. He has to guess the seat of the illness by 
amining the effect on the whole body. The physicist resembles a doctor, 
)t a watchmaker.* 

*P. Duhem, ThSotle physique; son objet—son structure (Paris, 1906), p. 24. 
‘Ibid., p. 285. 

*P. Duhem, "Quelques reflexions au sujet de la physique experimentale,” 
ivue des questions sdentifiques 36, 179 (1897). 

15 



modern science and Its philosophy 


t 

And elsewhere: » 

- The experimental verifications are not the basis of the theory, but its 
culmination. 

One notes how far Duhem has proceeded on the way from Mach’s con- 
ception of a physical theory to the conception which was later advo- 
cated by logical empiricism. 

Duhem, however, was also, from another angle, a great influence 
upon the philosophy of our group. He believed, like Mach, that an 
“explanation” that would be distinct from “economical description” 
would require an excursion into metaphysics. He says. 

If the object of physical theories is to explain experimental laws, physical 
theory is not an autonomous science; it is subordinated to metaphysics. 

Although he was convinced that physics cannot provide an explanation 
of experimental laws, he did not mean to say that no explanation is pos- 
sible. 

In seeking to stress the distinction between physics and metaphysics I do 
not mean to disdain either of these sciences, and I think to facilitate their 
accord much better than if I had confounded the object and the method of 
the former with the object and the method of the latter . . . The knowledge 
that metaphysics gives us of things is more intimate, more profound than 
that which is furnished by physics; it therefore surpasses the latter in 
excellence.^ 

As a matter of fact, Duhem was an advocate of Aristotelian and Thom- 
istic metaphysics and a faithful believer in the whole of Catholic the- 
ology. When I noted this merging of the most advanced type of “new 
positivism” with Thomistic metaphysics into one coherent system it im- 
pressed me strongly. Duhem interpreted on the basis of this hybrid 
philosophy the historic conflict between the Roman Church and the 
Copemican system. 

Although our group did not follow Duhem’s metaphysical predilec- 
tion, his doctrine became for us a frame of reference to which we could 

’’ P. Duhem, “Physique et metaphysique,” Revue des questions scientifiques 36, 
55 (1897). 


16 



historical background 


relate all 4he conflicts that have raged between science and religion 
and, more generally, between science and political ideologies. 


8. E. Mgch as a Philosopher of "Enlrghtenment'' 

Despite the aversion toward Mach that had been particularly notice- 
able among the German scientists and philosophers, he exerted a re- 
markable attraction for some groups of scientists as well as of philos- . 
ophers. However, the hostile reaction against him was of great intensity. 
Disputes about him have been repeated again and again at gatherings 
of scientists, the conflict of opinions has always been intense, and the 
discussions have frequently had an emotional flavor. 

Ernst Mach died in 1917, the year in which the Soviet government 
seized the power in Russia. Few among the European and American 
scientists realized at the time that the new ruler of Russia had pub- 
lished a book ten years previously in which he branded “Machism” 
as a reactionary philosophy. This book laid the foundation for a pe- 
culiar situation in the new Russia by making Mach a permanent target 
of attack. He was accused of agnosticism, subjectivism, and relativism. 
It was alleged that Mach had denied that science could know anything 
about the objective world. In this way, it seemed, the door was opened 
to other ways of finding the truth, particularly to the ways of traditional 
religion. It is interesting and amazing that Lenin denounced Machism 
by the same argument which has been used by advocates of physical 
realism among European and American scientists. 

In the year of Mach’s death (1917), 1 wrote Chapter 2 of the pres- 
ent book. Its purpose was the investigation of the value of Mach’s 
philosophy of science for the future development of science and of 
human thought in general. Mach’s philosophy is presented in such a 
way that those features are stressed that will survive him and remain 
essential parts of any future philosophy of science. This means in par- 
ticular that Mach is judged from the viewpoint of the “new positivism” 
of men like Poincar6 and Duhem. Mach’s philosophy is described and 
characterized with respect to its place in the history of human thought 
at the start of the twentieth century. It is likened to the philosophy of 
the Enlightenment in the eighteenth century. Mach analyzed the 


17 



modern science and its philosophy 


fundamental concepts of nineteenth-century physics, such as»mass and 
force, and made clear that all statements containing these words can be 
interpreted as statements about sense observations. Then concepts like 
these do not denote entities of a hidden real world behind the appear- 
ances, but are “auxiliary concepts” by which statements about observa- 
tions can be expressed in a more convenient and practical way. As the 
auxiliary concepts of nineteenth-century science were mainly concepts 
of mechanistic physics, Mach’s analysis was to a certain extent a de- 
bunking of mechanistic science as a system of statements about phys- 
ical reality. Nonetheless, Mach had no special bias against the mecha- 
nistic terminology that would imbue him with a particularly antimate- 
rialistic tendency. He tried to debunk all types of auxiliary concept 
in so far as they pretended to describe ontological realities or meta- 
physical entities. I illustrated this function of Mach’s philosophy as a 
philosophy of enlightenment appropriate to the turn of the century by 
pointing out similarities between Mach and Nietzsche. The great mass 
of writing about Nietzsche has largely overlooked the fact that he was 
a philosopher of enlightenment in his acute analysis of the auxiliary 
concepts of contemporary idealistic philosophy. 

The years after 1917 saw, as has been mentioned, the estab- 
lishment of Soviet power in Russia, the end of World War I, and the 
founding of new democracies in Central Europe, such as the Austrian, 
Czechoslovakian, and Polish republics. The event that had the greatest 
bearing at this time on the development of the philosophy of science, 
however, was the new general theory of relativity advanced by Einstein 
after 1916. In this theory Einstein derived his laws of motion and laws 
of the gravitational field from very general and abstract principles, the 
principles of equivalence and of relativity. His principles and laws were 
cormections between abstract symbols; the general space time co- 
ordinates and the ten potentials of the gravitational field. This theory 
seemed to be an excellent example of the way in which a scientific 
theory is built up according to the ideas of the new positivism. The 
symbolic or structural system is neatly developed and is sharply sepa- 
rated from the observational facts that are to be embraced. Then the 
system must be interpreted, and the prediction of facts that are ob- 


18 



historical background 


servable ipust be made and the predictions verified by observation. 
There were three specific observational facts that were predicted: the 
bending of light rays and the red shift of spectral lines in a gravitational 
field, and the advance of the perihelion of Mercury. 


9. A. Einstein's New Concept of Physical Science 

However, if we compare Einstein’s theory with previous physical 
theories, we note a certain difference in structure. It is after all only a 
difference in degree, but it directs our attention to a considerable 
change in the conception of physical theory. 

Whatever conclusions may be drawn from them, Einstein’s funda- 
mental laws will describe motions in terms of the general space-time 
coordinates. Before the results of his theory can be verified by observa- 
tion, it is necessary to know how statements about these general co- 
ordinates can be expressed in terms of observational facts. In traditional 
Newtonian physics, spatial coordinates and time intervals could be de- 
termined by the traditional methods of measuring length and time, by 
using yardsticks and clocks. However, the general coordinates in Ein- 
stein’s theory are quantities that define the positions and motions of 
moving particles with respect to systems of reference that can possess 
all sorts of deformations, with variable rates of deformation at every 
point. No rigid and defined system of reference for space and time 
measurement is given as a general basis for the definition of the space 
and time coordinates. The methods of measurement must be developed 
along with the ccnclusions from the principles of the theory. What is 
the bearing of these facts upon our general conception of the structure 
of a scientific theory? 

In nineteenth-century physics the translations of statements that 
contain abstract symbols of the theory— mass, distance, time interval, 
and the like— into observational facts did not cause much trouble. It 
was taken for granted that the straightness of a hne, the temperature of 
a body, or the velocity of a motion could be measured. At least, it was 
not suspected that there was any difficulty in assuming that such meas- 
urements are possible. In Einstein’s general theory of relativity, how- 
ever, the description of the operations by which these quantities could 


19 



modern science and its philosophy 


be measured becomes a serious and complex task; it becomeg an essen- 
tial part of the theory. These descriptions of the operations by which 
abstract symbols, such as the general space-time coordinates, are con- 
nected with observational facts are called today “operational defini- 
tions,” according to a terminology suggested by P. W. Bridgman.® 

As early as 1905, in his restricted theory of relativity, Einstein was 
well aware that the “operational definitions” are an essential part of his 
theory. Later he described the decisive alterations that were brought 
about by his new physical principles in the conception of a physical 
theory by stressing the fact that the connection between the symbols 
of the theory and the observational facts following from them is much 
longer, much more complex, and much more difficult to deal with than 
the connection assumed by nineteenth-century physics, to say nothing 
of the physics of the seventeenth and eighteenth centuries. The altera- 
tion brought about in the general conception of a scientific theory was 
a greater emphasis on the gap between the structural system and the 
experimental confirmation. Advancing a new theory now involved two 
tasks, both of which required great creative power: the invention of a 
structural system, and the working out of operational definitions for its 
symbols. 

However, the great new idea in every new physical theory, accord- 
ing to Einstein, was the creation of the structural system. In this sense 
a physical theory describes “the structure of the world.” This way of 
speaking could easily be interpreted as meaning that the symbols, 
which are the building stones of the structure, are also the “real build- 
ing stones” of the universe and that the structure of the symbolic sys- 
tem is “the real structure of the world." Following Einsteins ovra in- 
terpretation of his conceptions, statements like “the theory describes 
the real structure of the world” mean that appropriate operational defi- 
nitions enable us to derive from the symbolic system observational facts 
that check with our actual observations. Hence, the conception of phys- 
ics advocated by the new positivism of Poincar6 is altered by Einstein’s 
conception in such a way that a theory remains an economical descrip- 

®P. W. Bridgman, The Logic of Modem Physics (New York; Macmillan, 
1927 ). 


20 



historical background 


tion of facts by means of a structure and operational definitions.. How- 
ever, the connection between the symbols and the observational fact is 
not so simple as was anticipated. 


10. Geometry and Experience 

.The old problem of the relation between reasoning and experience 
in geometry was satisfactorily solved by Einstein’s theory? Even the 
advocates of a new positivism, like Poincare and Duhem, left a feeling 
of uncertainly concerning the actual position of geometry between the 
domains of logic and experience. One might even say that they left a 
feeling of uneasiness. J’oincare made it clear that geometry itself as a 
logical system says nothing about the physical world. From the view- 
point of physics, or of empirical science in general, this logical system 
can be judged only as a practical tool which can be used for a descrip- 
tion of the physical world. This procedure may be easy or difficult, 
simple or comphcated. By using Einstein’s theory, a description of mo- 
tions is obtained in terms of spatial and temporal coordinates that are 
measured by the reading of yardsticks and clocks. However, these in- 
struments, and therefore their readings, are affected by the same gravi- 
tational field that is responsible for the motion. An important conse- 
quence is that the distances between different points in space follow 
the axioms and propositions of Euclidean geometry only if the field of 
force is almost negligible. If we have to consider “strong” fields, the dis- 
tances between points no longer fit the rules of Euclidean geometry. 
This means, for instance, that the sum of the angles in a rectilinear tri- 
angle is no longer exactly equal to two right angles. The departure from 
two right angles is the greater the larger the area of the triangle and 
the stronger the field of gravitation. In order to give to these statements 
a physical meaning we assume tacitly that “straight lines” are defined 
by the traditional technologic procedures by which straight edges are 
produced on rigid bodies, such as a bar of steel^ 

This means in ordinary geometrical parlance that Euclidean geome- 
try is valid only if the field of gravitation is negligible and if the areas 
of triangles are small. Generally the laws that determine the relation 
between the measured distances will be those of non-Euclidean geome- 


21 



modern science ond its philosophy 


try. ^oincar^ and his immediate followers used to say that it is a con- 
vention whether one accepts Euclidean or non-Euclidean geometry. 
Actually this statement means only that we can build up structural 
systems of two types: Euclidean and non-Euclidean. As a structural or 
axiomatic system both are equally acceptable. If traditional methods 
of measurement are used— that is, if length and time are defined in 
terms of operations with rigid yardsticks and orthodox clocks— the fact 
must be considered that the instruments of measurement are affected 
by gravitational fields and therefore that the results of measurement, 
too, are affected. For every specific field of gravitation, the theory 
yields a specific influence on the yardsticks and clocks and therefore 
upon the results of measurement. By considering this influence “opera- 
tional definitions” can be formulated. The statements of mechanics 
now become statements about the results of actual measurements or 
about actual observations. They are therefore no longer arbitrary. If 
they are checked, they will be found to be either true or false. In a 
gravitational field the yardsticks and clocks will be affected in such 
a way that the results of measurements follow non-Euclidean rather 
than Euclidean geometr}^ 

In 1921, Einstein gave a lecture at the Academy of Sciences in 
Berhn with the title “Geometry and Experience,” ® in which he sum- 
marized in a conclusive way the place of reasoning and experience in 
geometry. Recording to Einstein, geometry can be either a structural 
system with arbitrary axioms or a physical theory. In the first case 
the conclusions of geometry are certain but do not tell us anything 
about the world of experience; in the second case the propositions of 
geometry can be checked by experiment and are as certain or uncertain 
as are any statements of physics or, for that matter, of any empirical 
science.] 

By Einstein’s argument it was demonstrated with great lucidity that 
there is no statement in geometry that is derived by reasoning without 
sense observations and that at the same time tells us something about 
the external world. Such geometric propositions, however, had been 

” An English translation of this lecture appeared in A. Einstein, Sidelights on 
Relatioity (London: Methuen, 1022). 


22 



historical background 


considered the most conspicuous examples of^ssertions about the ex- 
ternal world that are derived from pure reasoning. If the statements 
of geometry do not have this property, then the scientific basis of 
traditional metaphysics has disappeared. The cause of positivism 
against metaphysics has won a major battlq/ 

11. NeO'Kantianism and Neo«Thom!sm 

When Einstein had cleared up the foundations of geometry, the 
believers in scientific metaphysics were put on the spot; they were 
deprived of their best example of the existence of metaphysical asser- 
tions in science itself. 

Cam schools of traditional philosophy have more or less one belief in 
common: we can make general statements about facts of the external 
world with such certainty that they will never be refuted by any 
advance in science. For Aristotle such statements were the 
’ " of Greek astronomy, for Kant they were Newton’s principles 
Both the Aristotelians and the Kantians of the nineteenth 
and twentieth centuries agreed that the validity of Euclidean geome- 
try had been established forever. The only difference was that for one 
school the doctrine that was believed to prevail forever was Aristotle’s 
physics, while the other, more modem, school permitted the pursuance 
of the advance of science up to Newtot^ Each philosophical creed 
petrified the state of physics that prevailed at its time. 

^When Einstein demonstrated Ae possibility or even the plausibility 
that Euclidean geometry might be wrong he produced a catastrophic 
effect upon all the schools of traditional philosophy. The metaphysical 
schools of the Aristotelian and the Kantian types lost their basis in 
science,J^o meet this situation two attitudes have developed within 
these schools. To borrow a terminology from theology, we may call 
them the fundamentalist and the modernist attitudes. The first group 
maintain ed bluntly and boldly that scientists were just wrong. They 
were specialists and not able to reason correctly because of their lack 
of metaphysical training. 

The modernists are found in the camps of neo- Aristotelians (mostly 
called neo-Thomists) and of the neo-Kanti^s. They admitted that 


23 



modern science and its philosophy 


Euclidean geometry might be wrong. This was, according to these 
schools, the truth, but “not the whole truth." From the viewpoint of 
science proper, they have accepted fully the conception of geometry 
and physics advanced by the new positivism. But in the background of 
their teaching on geometry and physics there is a hint of a more pro- 
found wisdom which is presented in terms of Aristotelian or Kantian 
metaphysics.'^As an example of neo-Thomism we may consider the 
works of J. Maritain, particularly his book, “The Degrees of Knowl- 
edge.” An example of neo-Kantianism is the writing of Ernst Cassirer, 
particularly his books on the “Theory of Relativity” and on “Determi- 
nism and Indeterminism in Modem Physics.” The latter one is dis- 
cussed in Chapter 9 of the present book, ^his metaphysical back- 
ground, according to these authors, has no relevance for science proper; 
it is separated by airtight walls from the domain of scientific dis- 
course. In this way science became autonomous with respect to meta- 
physics, but the validity of the metaphysical assertions in the back- 
ground could not be checked by any experimental test. These assertions 
became more or less a tautological system of propositions, like pure 
mathematics or formal logic.“ However, this did not exclude the 
possibility that this metaphysical background may provide some satis- 
faction to readers or listeners.^s a matter of fact, the metaphysical dis- 
course has been usually couched in language that has evoked a pleasant 
resonance in people’s minds. This kind of language has been in use to 
present cherished types of science, ethics, or religion. If we disregard 
temporarily this background there is a large common ground between 
neo-Thomism, neo-Kantianism, and the new positivism. In extreme 
cases this common ground may be so extensive that one can read 
himdreds of pages of a neo-Thomist or a neo-Kantian without recog- 
nizing that he is not a positivist of the new type. The most outstanding 

“ J. Maritain, Distinguer pour unit, ou les degris du savoir (Paris: Brouwer, 
1935); English translation by B. Wall and M. R. Adamson (New York: Scribner, 
1938). 

Cassirer, Zur Einsteinschen relativitiUstheoric (Berlin: Cassirer, 1921). 

i*E. Cassirer, Determinismus und Indeterminismus in der modemen Physik 
. . . (Goteborg: Elanders, 1937). 

“C. W. Morris, Signs, .Language, and BehaviOT (New York: Frentice-Hall, 
1940), pp. 175 ff. 


24 



historical background 


exampte is the French physicist and philosopher, Pierre Duhem, whom 
we mentioned above. His writings are among the most valuable con- 
tributions of the new positivism. He was warmly recommended by 
Ernst Mach as a positivist. Many scientists were never aware that 
Duhem’s background was straight Aristotelian or rather Thomistic 
metaphysics. 


12. "New Wine into New Bottlei" 

The neo-Thomist and neo-Kantian schools reacted to the revolu- 
tionary changes that have arisen in science since the turn of the 
century by establishing a kind of “iron curtain” between science and 
philosophy. But none of these schools, and, as a matter of fact, none 
of the schools of traditional philosophy, of the idealistic or realistic 
type, were able to make a valuable contribution toward integrating 
the new science of the twentieth century into the general framework 
of human thought. From the viewpoint of intellectual history it is fair 
to say that the neo-Thomist and neo-Kantian schools have contributed, 
in a way, to the advance of scientific thought. They have helped to 
disintegrate the traditional systems to such a point that the remaining 
parts of the structure could easily merge with the new philosophy that 
would eventually arise on the basis of a new science. • 

[^Immediately after Einstein had published his general theory of 
relativity (1917), in which he advanced his new physics in full gen- 
erality, writings appeared that did not attempt to integrate the new 
physics into traditional philosophy but to build up a new philosophy 
on the basis of that new science^ These writings did not follow the 
reaction of traditional philosophy— of either the fundamentalist or the 
modernist types— to the new science. Theirs was a radical reaction in 
accordance with the words of the gospel: 

No man putteth new wine into old bottles; else tbe new wine will burst 
the bottles, and be spilled, and the bottles shall perish. But new wine must 
be put into new bottles; and both are preserved. No man also having drunk 
old wine straightway desireth new: for he saith. The old is better. 

Tbe old bottles were the patterns of traditional philosophy and the 
new wine was twentietii-century science. A ^up of men went in for 


25 



modern science and its philosophy 


new bottles. They did not borrow the framework of Thoiristic or 
Kantian metaphysics but they borrowed a pattern that had grown up 
in the soil of modern science, the pattern of the “new positivism.” 
While men like Poincare and Duhem had used this pattern for strictly 
domestic consumption, to clear up their own back yard, the foundations 
of science, the new men who emerged after 1917 ventured to build up 
a new philosophy that was expected to replace the traditional systems 
of the Aristotelian or Kantian type. 

The new movement started about the time when the first world 
war ended (1918). New democratic republics were established in 
Central Europe: Austria, Czechoslovakia, Poland, and the Weimar 
experiment in Germany. They offered a favorable soil for the evolu- 
tion of a scientific world conception. A similar situation seemed to arise 
in Russia after the overthrow of the Czarist regime (1917). It is inter- 
esting to note how the turn from the democratic start to the establish- 
ment of a new authoritarianism was accompanied by a turn from the 
philosophy of the new positivism to a philosophy which was nearer 
to the Aristotelian and Kantian tradition. 

The first peak of the Central European movement toward a scientific 
world conception was reached about 1920. We can characterize it by 

• three books; M. Schlick, “General Theory of Knowledge”” (1918); 

• H. Reichenbach, “Theory of Relativity and Cognition a priori” “ 

1^(1920); and L. Wittgenstein, Tractatus Logico-Philosophicus (1921).'“ 

The link between these books and Einstein’s theory is M. Schlick’s 
small book "Space and Time in Contemporary Physics” (1917),” in 
which the author attempts an integration of the new positivism with 
the ideas that have grown out of Einstein’s new science. 

”M. Schlick, Allgemeine Erkenntnislehre (Berlin: Springer, 1918; ed. 2, 
1925). 

H. Reichenbach, Relatlvitdtstheorie und Erkenntnis apriori (Berlin: Springer, 
1920). 

'*L, Wittgenstein, Tractatus hgico-philosophicus (London: Paul, Trench, 
Trubner; New York: Harcourt, Brace, 1922); German and English on opposite 
pages; 

M. Schlick, Raum und Zeit in der gegenwartigen Fhysik, . . . (Berlin: 
Springer, ed. 3, 1920); English translation by H. L. Brose (Oxford: Clarendon 
Press; New York: Oxford University Press, 1920). 


26 



historical background 


13. M. Schlick's New Criterion of Truth 

The views of the new positivism just before Schlick entered the 
picture are presented in Chapter 2, which was written in the same 
year as Schlick’s “Space and Time.” In this chapter la distinction is 
made between observational concepts— red, warm, . . .—and auxiliary 
concepts— force, electric charge, ... All statements of science that 
contain auxiliary concepts are translatable into statements containing 
observational concepts only. The expression “auxiliary concept” was 
often interpreted as meaning that by words like “force,” electric 
charge,” “potential,” no “element of reality” is denoted. Schlick, how- 
ever, stresses that the usefulness of a statement in science depends 
only upon whether this statement can be checked by sense observa- 
tions. In other words, a scientific theory must consist of such principles"^ 
that statements containing observational concepts only can be logically 
derived from them. It is irrelevant whether the concepts in the theory 
itself are “observational concepts” or “auxiliary concepts.” 

Schlick says: 

There is no argument whatsoever to force us to state that only the in- 
tuitional elements [i.e., observational concepts], colors, tones, etc., exist in 
the world. We might just as well assume that elements or qualities which 
cannot be directly experienced also exist. These can likewise be termed 
“real,” whether they be comparable with intuitional ones or not. For ex- 
ample, electric forces can just as well signify elements of reality as colors and 
tones. They are measurable, and there is no reason why epistemology should 
reject the criterion for reality which is used in physics.^® 

To say that electric forces are “measurable” means that from statements 
of the form, “At this point of space acts an electric force of one hundred 
grams,” statements can be derived about the deviation of a pointer 
from its coincidence with a certain mark of a scale. This means, how- 
ever, statements about direct sense observation. It does not matter 
whether the words in a statement denote observational concepts or 
auxiliary concepts provided that results can be derived that contain 
only observational (intuitional) terms. 

“ffifcl., p. 84. 


27 



modern science and its philosophy 


For Otherwise no measurement would be possible. Therefore, the 
“operational definitions”— which link auxiliary and observational con- 
cepts— are an indispensable part of every theory. This point is very 
important for integrating Einstein’s new science into the scheme 
of new positivism. Now, one can, without any trouble, introduce terms 
like “four-dimensional space” or “curved space” into physics. These 
concepts are as legitimate as three-dimensional space and Euclidean 
geometry. We have only to make sure that our systems contain the 
rules by which, from the facts of the theory, only observational con- 
cepts can be derived. 

From this characterization of scientific theories Schlick proceeds 
to an even more radical departure from traditional philosophy. He 
starts from the method that is generally used in science to check the 
“truth” of a theory. One considers, for example, the electric charge of 
an electron, which, according to the theory, should have a constant 
numerical value. Then one derives from the theory by a chain of 
conclusions the result that “the charge of the electron is 4.7 X 10~’“ 
electrostatic units.” Then one tries to find, on the basis of the same 
theory, a second chain of conclusions by which the numerical value 
of the charge of the electron can be derived. If this value is different 
from the first one, we say that our theory is “not true.” If we obtain 
the same value as by the first method, we say that our theory may be 
true or “is confirmed.” The "degree of confirmation” is the higher the 
more independent chains we can find that lead to the same numerical 
value for the charge and for other concepts of the theory. 

This criterion of truth can also be formulated in a more general 
way. The symbols of the theory (observational or auxiliary concepts) 
are the state variables by means of which the course of events in the 
physical world— the facts of the world— can be described. Every specific 
state of the world is defined by specific numerical values which are 
assigned to these variables. Some of them, like the charge of the 
electron, are constants. Others, like the coordinates of an electron, are 
variables. As we learned by the example of the charge of the electron, 
we can make use of the theory to calculate these numerical values. 
This means that the theory indicates the facts of the world. If we 


28 



hstorical background 


obtained from two difiFerent chains of conclusions different numerical 
values for a symbol which, according to the theory, should be constant, 
then one and the same theory would indicate two incompatible facts. 
Such a theory would be useless. 

From these remarks one will understand fairly well Schlick’s 
criterion for the “truth” of a theory. He says: 

/Every theory is composed of a network of conceptions and judgments, 
and is correct or true if the system of judgments indicates the world of facts 
uniquely}^ 

If a theory yields two different values for the charge of the electron, the 
correspondence between the theory (concepts and judgments) and 
the facts would not be a unique correspondence. 

On the other hand, no harm is done if different theories indicate one 
and the same world of facts. In the direction from the facts to the 
theory no uniqueness is required of “true” theory. Schlick says: 

It is, however, possible to indicate identically the same set of facts by 
means of various systems of judgments; and consequently there can be 
various theories in which the criterion of truth [unique correspondence] is 
equally well satisfied . . . They are merely different systems of symbols, 
which are allocated to the same objective reality.™ 


14. M. Schlick and H. Relchenbach 

From this analysis of scientific theories Schlick proceeded to the 
claim that every cognition, in whatever domain of knowledge, is essen- 
tially the establishment of a correspondence between the facts of the 
world and a system of symbols. Since between these symbols a set of 
relations— for example, the axioms of geometry or mechanics— is estab- 
lished, an arbitrary correspondence would frequently assign several 
worlds of facts to the same set of symbols. Then the cognition is false. 
According to Schlick, a cognition is “true” if the correspondence es- 
tablished is unique. A system of symbols indicates the facts of the 
world uniquely. 


“Ibid., p. 86. 

“ Loc. cit. 


29 



modern science and its philosophy 


This conception of cognition and truth was a radical bref.k with 
almost all systems of traditional philosophy, according to which cog- 
nition meant the finding of a truth that was hidden behind the appear- 
ances and could be discovered there by the power of reason, which 
the trained philosopher was supposed to possess. According to Schlick, 
however, cognition is the establishment of a correspondence; this 
means, primarily, building up a system of symbols with relations 
among them. Cognition becomes an activity, the construction of a sys- 
tem of symbols that has only to fulfill the requirement of uniqueness. 
If we start from this conception of cognition, most of the metaphysical 
problems that have puzzled generations of philosophers lose their in- 
soluble aspect. 

Consider, for example, the ancient puzzle whether the world is es- 
sentially mind or matter, the answer to which has been the shibboleth 
for distinguishing between idealists and materialists. Schlick would 
reformulate the problem: Are mental concepts or physical concepts 
better suited to build up a system of symbols that could indicate the 
world of facts uniquely? The problem loses its yes-or-no character and 
becomes a problem of deciding, on grounds of convenience, between 
two ways of describing the facts of the world. 

Another author who ventured at that time to build up, on the basis 
of the “new positivism” and Einsteins new theories, a philosophy of 
science that would replace the traditional schools was Hans Reichen- 
bach. His most significant books in this line, which to a certain extent 
paralleled Schlick’s writings, are the “Theory of Relativity and Cog- 
nition a priori” “ and “Axiomatics of the Relativistic Theory of Space 
and Time.” ” Reichenbach went in much more for the technical dis- 
cussion of physical and philosophical problems than Schlick did. He 
approved Schlick s conception of “true cognition,” but he presented it 
in a way that was closer to the conception of science advanced by men 
like Poincare and Duhem. Reichenbach was perhaps the first one who 
formulated explicitly the requirement that every theory in which non- 
observational concepts appear contain relations between these abstract 

” H. Reichenbach, AxiomaUk der relativistUchen RaurtirZett-lehre ( Braun- 
schweig: Vieweg, 1924). 


30 



historical background 


s 

concepts and observational concepts. He stressed that geometry as an 
empirical science must contain, besides the geometric axioms (relations 
among geometric concepts), the description of a method whereby the 
straightness of a line can be tested experimentally. As the axioms of 
geometry and the description of measmrements form a net, we can re- 
gard the whole theory as a prescription for coordinating the abstract 
concepts of geometry with observational facts. In this sense, Reichen- 
bach regarded geometry as a system of “axioms of coordination.” His 
criterion of truth was then similar to Schlick’s, namely, that the axioms 
of coordination should be unique. 

This uniqueness could, as he stressed, be correctly decided only by 
observation. On this basis, Reichenbach discussed the question whether 
the new philosophy was compatible with the traditional Kantian sys- 
tem. Ry Einstein’s new science it was demonstrated that the Euclidean 
axioms of coordination may be false. Einstein described definite ex- 
periments by which one can test whether the traditional axioms of co- 
ordination lead to contradictions. Kant believed that the imiqueness of 
Euclid’s system of geometry is founded in the organization of the hu- 
man mind, while Reichenbach stressed that it has to be checked by 
experience. In the discussion (Secs. S and 4) of the relation of Poin- 
care to Kant I made partial use of Reichenbach’s views. 


15. L. Wittgenstein and R. Carnap 

The work done by Schlick and Reichenbach made a strong impres- 
sion upon our Viennese group (Sec. 1). In our search for a scientifically 
founded philosophy we were glad to find collaborators who attacked 
the task, I would say, from a more professional angle. At that time 
(after 1920) Hans Hahn was professor of mathematics at the Uni- 
versity of Vienna, Otto Nemath started working for the City of Vienna, 
organizing adult education in the social sciences, and I had been since 
1912 professor of theoretical physics at the University of Prague in 
the new Czechoslovakian Republic. Hahn had started intensive work 
with advanced students in the field of symbolic logic and the founda- 
tions of mathematics. In 1922, he chose as a basis of their discussions 
the new book by L. Wittgenstein, Tractatus^ Logico-Pkilosophicus,^ 


31 



modern science and its philosophy 


These discussions were the germ of many future developmenfs in the 
philosophy of science. 

Wittgenstein’s book followed in some respects the same line as 
Schlick’s and Reichenbach’s. Wittgenstein, who was also of Viennese 
origin, was a student of Bertrand Russell, the British logician and 
philosopher. He added a new and important component to the integra- 
tion performed by Schlick. Without making much use of the tech- 
nicalities of symbolic logic, Wittgenstein showed that the new philos- 
ophy could be brought into a more perfect and coherent shape with the 
aid of the basic ideas of Russell’s logic. Wittgenstein’s formulations 
sounded even more straightforward and provoking than Schlick’s, 
Reichenbach’s, and even Russell’s. Wittgenstein claimed bluntly that 
I the problems of traditional philosophy are merely verbal problems. 
Our ordinary language, which has grown up to describe the facts of 
everyday life, is not adapted to the task of expressing and answering 
problems put to traditional philosophy. If we try to use our ordinary 
language in this way, we get into trouble. The real problem is to find 
out what one actually can say clearly. The world of facts can be de- 
scribed in our ordinary language; therefore, says Wittgenstein, “to 
understand a proposition means to know what is the case if it is true.” 

This conception of “meaning” and “understanding” is essentially no 
different from Schlick’s, Reichenbach’s, or, as a matter of fact, Mach’s, 
if we understand Mach as he is interpreted in Chapter 2 of the present 
book. Every student of philosophy will also remember C. S. Peirce’ s 
conception of “meaning,” and even William James’s, who said that the 
meaning of a sentence is its "cash value .” 

Wittgenstein’s merit, however, was his precise logical formulation 
and his cutting and striking dialectic. His line was later called, quite 
adequately, "therapeutical positiv ism.” Hahn became very enthusiastic, 
starting a close cooperation of the new men with our Viennese group. 
He envisaged the appointment of M. Schlick as a professor of philos- 
ophy at the University of Vienna. He met, of course, a stiff resistance 
among the adherents of traditional philosophy. But the interest of the 
scientists in the philosophical background of science has been tradi- 
tionally high at the University of Vienna. Ernst Mach had owed his 

I 

32 



historical background 


appoinftnent to this predilection and Hahn succeeded in enlisting a 
sufficient number of scientists in a drive to carry through Schlick’s ap- 
pointment in 1922. In this year a close cooperation between Schlick 
and the old Vienna group began. This common work gained a great 
deal in intensity and momentum when Schlick persuaded R. Carnap 
to move to Vienna in 1926. 

Camap gave the new philosophy its “classical” shape. He coined 
many of its terms and phrases and endowed it with subtlety and sim- 
plicity. In the form created by Camap it became a center of interest 
and a target of attack on a large scale. 

Schlick and Reichenbach had identified “tme cognition” with a 
system of symbols that indicated the world of facts uniquely. Camap 
offered an example of such a system in his book “The Logical Stmcture 
of the World.” In this book the integration of Mach and Poincare 
was actually performed in a coherent system of conspicuous logical 
simplicity. Our Viennese group saw in Carnap’s work the synthesis 
that we had advocated for many years. 

Camap introduced as the elementary concepts of his system im- 
mediate sense impressions and the relations of similarity and diversity 
between them. The world is to be described by statements that may . 
contain any symbols, provided that from them statements can be 
logically derived that contain nothing but assertions about similarity 
between sense impressions. The “meaning” of a statement in science 
would be the sum of all statements about similarity and diversity be- 
tween sense impressions that can be derived logically from the state- 
ment in question. When I read this book it reminded me strongly of 
William James’s pragmatic requirement, that the meaning of any state- 
ment is given by its “cash value,” that is, by what it means as a direction*^ 
for human behavior. I wrote immediately to Camap, "What you advo- 
cate is pragmatism.” This was as astonishing to him as it had been to 
me. We noticed that our group, which lived in an environment of 
idealistic philosophy had eventually reached conclusions by which we 
could find kindred spirits beyond the Atlantic in the United States. 

From Carnap’s presentation it was clear that a system from which 

R. Camap, Der logische Aufbau der WeU (Berlin: Weltkreis-Verlag, 1928). 

33 



modern science and its philosophy 


one cannot derive results about similarities between sense imj^essions 
caimot be a “true cognition.” The statements of traditional metaphysics, 
such as those about the existence or nonexistence of the external world, 
can obviously not be statements of the required type. For this reason 
Carnap said that “metaphysics is meaningless.” 


16. O. Neurath's "Index of Prohibited Words" 

The men who had expanded the new positivism into a general log- 
ical basis of human thought— Schlick and Carnap— came now into per- 
sonal contact with the original Viennese group, particularly with Hahn 
and Neurath, while my own contact was restricted to the time of the 
university vacations. As a result of the developing cooperation the 
new philosophy became more and more different from the traditional 
German philosophy, to which both Schlick and Carnap were bound 
to have some sentimental ties originating from their training at German 
universities. 

They had demonstrated logically that no scientific metaphysics is 
possible because metaphysical statements do not fit into the pattern 
that statements must have in order to be called true or false. But the 
social scientist Neurath investigated the meaning of metaphysical state- 
ments as social phenomena. He insisted with a certain nithlessness that 
no formulation should be allowed to slip in that would “give comfort 
to the enemy,” even if it would be admissible from the purely logical 
viewpoint. The whole original Viennese group was convinced that 
the elimination of metaphysics not only was a question of a better 
logic but was of great relevance for the social and cultural life. They 
were also convinced that the elimination of metaphysics would deprive 
the groups that we call today totalitarian of their scientific and philo- 
sophic basis and would lay bare the fact that these groups are actually 
fighting for special interests of some kind. 

In the long thorough and intimate discussions that the Viennese 
group had with Schlick and Carnap, the point was made that the new 
philosophy must be built up in such a way that no misinterpretation 
in favor of metaphysics could occur. We all knew that misinterpreta- 


34 



historical background 


tions were bound to happen if and when expressions like “real,” “essen- 
tial,” “real building stone of the universe,” were used in a loose way. 
Neurath even recommended, half jokingly, that an “index of prohibited 
words” should be set up. In a monograph “ on the “Foundations of 
the Social Sciences” he avoided, as he explicitly states, words like “en- 
tity,” “essence,” “mind,” “matter,” “reality,” “thing.” 

A well-known example of misinterpretation of this type is the praise 
or condemnation of Machs philosophy as a brand of idealism be- 
cause his doctrine was often presented as claiming that the “world 
consists actually only of sensations.” This was interpreted again as 
meaning that the “world is essentially mental.” This interpretation ac- 
counts for Lenin’s violent attack on Mach and for the extremely hostile 
attitude of the official Soviet philosophy against all doctrines that 
traced their origin back to Mach. This holds also for the new positivism 
and its generalization as achieved by Schlick, Reichenbach, and Car- 
nap. 

Perhaps the most striking effect of the cooperation of Schlick 
and Carnap with the old Viennese group was the shift to “physicalism” 
and to the “unity of science.” Neurath had been particularly eager to 
prohibit any establishment of a metaphysical doctrine by a tactic of 
infiltration. He suggested that sense data should be dropped as ele- 
mentary concepts of the logical structure of the world and replaced 
by physical things. Instead of building up the system of human knowl- 
edge upon concepts like "red spot” or "feeling of warmth,” one should 
use elementary symbols expressing concepts like “rock” or “table,” 
and define “redness,” or “warmth” as derived concepts. As the starting 
point in sensation had a certain tint of idealism, so the new starting 
point had a tint of materialism. Carnap had in his “Logical Structure 
of the World” spoken of “methodical materialism” as a possible lan- 
guage for his system. But he had come to prefer “phenomenal lan- 
guage,” statements in terms of sense impressions. Neurath worked out 
a system based on physical things as elementary concepts and called by 

^International Encyclopedia of Unified Science (Chicago; University of 
Chicago Press, 1938-39), vol. 2, no. 1. 


35 



modern science and its philosophy 


him “physicalism.” Camap refined Neurath’s physicalism to a®precise 
logical structure and even constructed a “physicalistic language” for 
the field of psychology. 

This transition from a quasi-idealistic to a quasi-materialistic lan- 
guage, which took place in our group about 1930, has been misunder- 
stood by great many authors. They interpret it as a sudden jump into an 
opposite type of philosophy. As a matter of fact, the “jump” was an 
expression of our firm belief that the difference between an idealistic 
and a materialistic system is logically and scientifically of little im- 
portance and that there is actually only a difference of emphasis. The 
choice is determined largely by the emotional connotations or, in other 
words, by the language in which the pattern of our general culture 
is usually described. 


17« O. Neuroth and the "Unity of Science Movement" 

The second reformulation suggested by Neurath was the char- 
acterization of the new movement as a work “towards the unification 
of science” or “for the construction of a unified science.” This shift 
was for many people surprising too. Schlick and Carnap had stressed 
the point that there are no philosophic propositions but that there is a 
philosophic activity that consists in the clarification of the statements 
of the special sciences. This meant, briefly, that philosophy was to 
interpret the abstract and symbolic principles of science as state- 
ments about physical things. Frequently, particularly in Great Britain, 
the new philosophy has been distinguished from the traditional phi- 
losophy as “analytic philosophy” in contrast to “speculative philosophy.” 
The work of men like Moore, Russell, or Wittgenstein has been de- 
scribed as analytic philosophy. 

Our original Viennese group and particularly Neurath were not 
satisfied with ascribing to our new philosophical group mainly critical 
and analytical objectives. We knew well that man is longing for a 
philosophy of integration. If the new philosophy refuses to serve the 
cause of integration, a great many people, including even scientists, 
would rather return to traditional metaphysics than be restricted to a 
purely analytic attitude. As a matter of fact, the traditional goal of 


36 



historical background 

« 

“philosephy,” through thousands of years of human knowledge, has 
been integration. 

Neurath pointed out that by Carnap’s analysis the statements of 
all sciences, not only physics but also biology and sociology, had been 
reduced to statements about physical things or about sense impressions. 
The traditional opinion about the individual sciences has, however, 
been that the fundamental concepts of biology are essentially different 
from those of physics, the concepts of psychology different from those 
of biology and so on. Physics, biology, and psychology have to do with 
different kinds df “being” and can never be united on the level of 
science. Only by introducing metaphysical concepts can one achieve 
a unification. If we accept, however, Carnap’s analysis of science, it 
follows that all statements of science are of only one type, that is, they 
are statements that can be expressed in the “thing language.” Hence, it 
must be possible to introduce a unified language for all the sciences and 
to create a system of "unified science,” in which the “special sciences” 
are merely products of the division of labor. The terminology of the 
special sciences is practical for restricted purposes, but no philosophic 
implication about unbridgeable gaps can be drawn from the differences 
in terminology. 

The new philosophy now described its work as the building up of a 
unified science. With this goal we returned in some measure to the 
classical goal of philosophy as defined by Aristotle. As a matter of fact, 
August Comte, the father of positivism, said (1829): 

I employ the word “philosophy” in the sense that was given to it by 
Aristotle, as denoting the general system of human conceptions. 

Our group did not wish to stress die work on analysis in contrast to 
the creation of a synthesis. We never regarded the logic and analysis 
of science as a goal in itself; we believed strongly that this analysis 
is a necessary part of obtaining an unprejudiced ^outlook on life. The 
close connection between the positivistic attitude and the unification 
of science can be traced back to Ernst Mach himself. In Chapter 3 
of the present book I analyze Mach’s philosophy from the viewpoint 
of 1938, the one-hundredth anniversary of his birth, and point out to 


37 



modern science and its philosophy 

what degree the further evolution of positivism was anticiQ^ated in 
Mach’s writings. 


18. The Vienna Circle 

In 1929, we had the feeling that from the cooperation that was 
centered in Vienna a definite new type of philosophy had emerged. As 
every father likes to show photographs of his baby, we were looking 
for means of communication. We wanted to present our brain child 
to the world at large, to find out its reaction, and to receive new 
stimulation. 

We decided first to publish a monograph about our movement, 
next, to arrange a meeting, and eventually to get control of a philo- 
sophical journal so that we would have a way of getting the contri- 
butions of our group printed. 

When we prepared the monograph we noticed that our group and 
our philosophy had no name. Quite a few people among us disliked 
the words "philosophy” and "positivism” and id not want them to 
appear in the title. Some disliked all “isms,” foreign or domestic. 
Eventually we chose the name “scientific world conception.” Some 
of us, particularly Schlick, thought that every reasonable scientist 
would agree with our presentation of cognition. Our chosen title 
seemed a little dry to Neurath, and he suggested adding “The Vienna 
Circle,” because he thought that this name would be reminiscent of 
the Viennese waltz, the Vienna woods, and other things on the pleas- 
ant side of life. The monograph ’’ was written by Carnap, Hahn, and 
Neurath in close cooperation. 

Two years later, A. Blumberg and H. Feigl published in the United 
States a paper, “Logical positivism; A new movement in European 
philosophy,” “ and provided the “scientific world conception” with its 
international trade name. 

“We chose the term “world conception” (Weltauffassung) in order to avoid 
the German word Weltanschauung, which seemed to us loaded with metaphysical 
connotations. 

Wissenschaftliche Weltauffassung der Wiener Kreis (Wien: A. Wolf, 1929). 

“A. Blumberg and H. Feigl, Journal of Philosophy 28, 281 (1931). 


38 



historical background 


In accord with the historical and cultural inclination of the Vien- 
nese group, Neurath was eager to trace the genealogic lineage of our 
movement. In the monograph he recorded the following lines: 

Positivism and empiricism; Hume, the philosophers of the En- 
lightenment, Comte, Mill, Avenarius, Mach. 

Scientific method: Helmholtz, Riemann, Mach, Foincar^, Enriques, 
Duhem, Boltzmann, Einstein. 

Symbolic logic and its application to reality: Leibniz, Peano, Frege, 
Schroeder, Russell, Whitehead, Wittgenstein. 

Eudaemonistic ethics and positivistic sociology: Epicurus, Hume, 
Bentham, Mill, Comte, Feuerbach, Marx, Spencer, Mueller-Lyer, 
Popper-Lynkeus, Carl Menger. 

A great many misunderstandings have been current about the orig- 
inal doctrine of the Vienna Circle, which became the germ of logical 
positivism. Again and again philosophers have sought to prove to the 
“positivists” that there is a “real world” and that science explores this 
real world. In fact, the monograph says: "Something is real if it is a 
part of the system of symbols that denotes the world of facts.” A num- 
ber of authors have tried to refute “positivism” by showing that there 
are also unobservable elements in science that have to be called “real," 
for example, the electron. The quoted passage clearly expresses the 
fact that “real” is not regarded as identical with “observable.” We can 
state the meaning of the quotation also in slightly different words: 
Every symbol denotes something real if this symbol is a part of a^' 
system that serves to describe observable facts uniquely. 


19. The First Public Meeting 

The arrangement of the meeting was not so easy. We wanted to 
reach a large audience. The ordinary regular philosophy meetings 
followed the traditional lines and would hardly have given us enough 
scope. By a happy coincidence I was just in 1929 arranging a meet- 
ing of the physicists and mathematicians from the German-speaking 
regions in Central Europe. The meeting was to be held in my place 
of residence, Prague, the capital of Czechoslovakia. The German 


39 



modern science and its philosophy 


Physical Society, which was the oiEcial sponsor of this meeting, did 
not particularly like the idea of combining this serious scientific meet- 
ing with such a foolish thing as philosophy. However, I was the chair- 
man of the local committee in Prague and they could not refuse my 
serious wish to attach a meeting with the topic, “Epistemology of the 
Exact Sciences.” This meeting was to be sponsored by the Ernst Mach 
Association, which was the legal organization of the Vienna Circle, 
and the Society for Empirical Philosophy, which was organized in 
Berlin and followed in general the line of H. Reichenbach. In this way 
we provided a nucleus of interested people and hoped that quite a 
few mathematicians and physicists who came to Prague for their 
meeting would also attend our gathering. 

Some scientists wanted to minimize our program and predicted 
that we would have no audience from the ranks of the exact scientists. 
As a matter of fact, our addresses had a larger audience than papers 
on special scientific problems. I had prepared an elaborate paper that 
was intended to give the scientists a kind of preview of our ideas and 
to prove that the new line in philosophy is the necessary result of the 
new trends in physics, particularly the theory of relativity and the 
quantum theory. I elaborated the contrast between the “school philos- 
ophy,” which sought to preserve the traditional doctrines despite the 
new science, and the “scientific world conception,” which wants to 
pour the new wine into new bottles. I stressed also the similarity be- 
tween the new scientific world conception and the basic ideas of 
American pragmatism, particularly those of William James. This lec- 
ture is Chapter 4 of the present book. 

Some friends cautioned me not to speak too bluntly. The audience, 
which consisted mostly of German scientists, knew little of philosophy, 
except that they had some sentimental ties to Kantianism. This doc- 
trine was regarded in some intellectual quarters as a kind of .substi- 
tute for the traditional forms of religion. My wife said to me after the 
lecture: ‘Tt was weird to listen. It seemed to me as if the words fell 
into the audience like drops into a well so deep that one cannot hear 
the drops striking bottom. Everything seemed to vanish without a 
trace.” 


40 



historical background 


Ther# is no doubt that quite a few people in the audience were 
shocked hy my blunt statements that modem science is incompatible 
with the traditional systems of philosophy. Probably, most of the 
scientists had not been accustomed to thinkin g of philosophy and 
science as one coherent system of thought. Philosophy had been for 
them what the Sunday sermon is for a businessman who is only inter- 
ested in profit. Philosophy had been required not to be “tme” but to 
give emotional satisfaction. 

After the meeting, however, our committee received a great many 
letters from scientists who expressed their great satisfaction that an 
attempt has been made toward a coherent world conception without ’ 
contradictions between science and philosophy. We even received 
one letter from a professor of philosophy at a German university who 
wanted to go on record with his conviction that the remaking of 
philosophy, along the lines that we followed at our meeting, is 
necessary. 

20. M. Schlick Announces a 'Turn in Philosophy" 

In the years 1930-31, tliere appeared the first volume of the jour- 
nal Erkenntnis (Cognition), which became the main mouthpiece of 
our movement. The editors were R. Caniap and H. Reichenbach. The 
first issue began with the paper, “The Turn in Philosophy," by 
M. Schlick. I shall quote some lines in order to show that an opti- 
mistic belief in the new trend was the keynote of this journal. Schlick 
writes: 

I am convinced that we are in the middle of an altogether final turn in 
philosophy. I am justified, on good grounds, in regarding the sterile conflict 
of systems as settled. Our time, so I claim, possesses already the methods by 
which any conflict of this kind is rendered superfluous; what matters is only 
to apply these methods resolutely. 

In the same year (1930) Schlick published a paper, “Personal Ex- 
perience, Cognition, Metaphysics,”” in which he writes: 

All cognition of the being is achieved, in principle, by the methods of die 
special sciences; every other kind of ontology is empty talk. Metaphysics is 

” M. Schlick, “Erleben, Erkennen, Metaphysik," Kantstudien 31, 146 (1926). 

41 



modern science and its philosophy 


impossible because its goals contradict one another. If the metaphysician 
longs only for personal experience, his longing can be satisfied by poetry 
and art— or by life itself. But if he longs for a personal experience of the 
transcendent, he confuses life and cognition, he chases futile shadows. 

For Schlick, as we know, “cognition” is the construction of a system of 
symbols that denotes uniquely the world of facts. It is therefore funda- 
mentally different from personal experience. 

This strong optimistic feeling is psychologically the feeling of a 
turn. You can ride in a car at high speed and you do not feel any- 
thing so long as the velocity remains unchanged. But if a turn or an 
acceleration takes place, you experience a strong reaction. Today, the 
movement of logical positivism is no longer so conspicuous. It had 
produced a turn in philosophy, which afterwards moved in a new 
direction and rather smoothly. I quote a passage from a philosopher 
who is by no means a follower of what is now called logical positivism. 
C. West Churchman writes: “ 

Few can doubt the healthy impact that the positivist position has had 
upon modes of inquiry; it has sharply distinguished the schools of thought, 
and has raised a standard under which the proponents of experimental 
methods can fight their battles against a reactionary movement. To return to 
a prepositivistic viewpoint is to retiun to a prescientific viewpoint, to become 
a reactionary as an advocate of the indisputable power of the sovereign in 
the eyes of one with a democratic outlook. 

We find a similar position even in the most recent book of F. S. C. 
Northrop “ who, in some sense, attempts a justification of metaphysics. 
He has been, however, in all his writing a very independent thinker 
who has given much thought to the foundations of science and par- 
ticularly to the interdependence between the philosophy of science 
and die cultural background. He writes: 

“In any event, the great merit of logical positivism and its main aim is 
satisfied, even if one leaves the scientific concepts and their meaning just as 

® C. W. Churchman, Theory of Experimental Inference (New York: Macmil- 
lan, 1948). 

*®F. S. C. Northrop, The Logic of the Sciences and the Humanities (New 
York: Macmillan, 1947), pp. 113, 114. 


42 



historical background 


one finds^them, as prescribed by the scientists in the postulates of some 
specific deductively formulated theory. The important desideratum at which 
the logical positivists were aiming, namely operational verification, can 
nonetheless be obtained. There are many signs that contemporary logical 
positivists have now come to this position. 

As a matter of fact, if one traces the history of logical positivism, one 
will see that this has been always the position of the “scientific world 
conception.” This becomes clear if one considers Schlick’s position in 
“Space and Time” (Sec. 13). 

To estimate how sharp the turn was for which logical positivism 
was responsible, we have to compare its position with the views of the 
school of traditional philosophy that was the nearest to it in spirit and 
in time. We choose for comparison H. Vaihinger’s “Philosophy of ‘As 
If”’ (1911), a very ingenious and in its time very famous book, a typi- 
cal example of what I called the disintegration of traditional philosophy 
by neo-Kantianism. Vaihinger tries to show that the concept of an 
atom (which he identifies with a mass point) in physics is a useful 
"fiction” although it is logically self-contradictory. He says: 

An entity without extension that is at the same time a substantial bearer 
of forces— this is simply a combination of words with which no definite mean- 
ing can be connected. “Simple atoms,” that must yet be something material, 
cannot be causae verae, cannot be actual things. Since, however, the physi- 
cist does require atoms for his construction, how is this contradiction to be 
solved? How are we to rescue science from this dilemma? “ 

Vaihinger thinks that the method actually used in science is to 

speak of atoms without really meaning to assume them . . . Unquestionably 
this conceptual method is die most convenient one, but this constitutes, of 
course, no proof of its objective-metaphysical validity." 

Vaihinger’s view is a clear indication of the situation in philosophy 
immediately before the rise of logical positivism. There was a com- 

“Hans Vaihinger, Die PhUosophie der “Ah Ob" (1911); English translation 
by C. K. Ogden (ed. 2, Bames and Noble, 1935), p. 219. Incidentally, in this 
book the term “logical positivism” was used for the first time, although in a some- 
what different sense from the one it now has. 

»16id.,p. 222. 


43 



modern science and Its philosophy 


plete lack of understanding that one must distinguish between a 
structural system having exact logical coherence with the world of 
facts, which are described with a certain vagueness, and the opera- 
tional definitions, which connect both domains and participate in the 
preciseness of the first and the vagueness of the second. 


2 \» P. W. Bridgman's Theory of Meaning 

About the same time when the “scientific world conception” group 
were arranging their first public meeting, P. W. Bridgman published a 
book ® in the United States in which he reacted to the same situa- 
tion by which this group had been faced. In a broad sense, we can 
characterize his work also as an attempt to integrate Mach, Poincare, 
and Einstein into a coherent picture of modern science. Bridgman’s 
field was not mathematics or symbolic logic but experimental physics. 
He has been a man of the laboratory who preferred to do things 
rather than set up a long chain of arguments. His approach is therefore 
different from that of the Central European group. It was, in a way, 
more similar to Mach’s, who was also essentially an experimentalist. 
Bridgman found out what was the salient point in the integration of 
Mach and Poincare. Reichenbach had explicitly pointed out that 
what is needed is a bridge between the symbolic system of axioms and 
the protocols of the laboratory. But the nature of this bridge had been 
only vaguely described. Bridgman was the first who said precisely 
that these “relations of coordination” consist in the description of 
physical operations. He called them, therefore, “operational defini- 
tions.” This name has been generally accepted. Bridgman was also 
very definite in stating that a theory which does not contain the opera- 
tional definitions of its abstract terms is meaningless. In this way he 
arrived at a concept of “meaning” and “meaninglessness” that was 
similar to the concept advanced by Carnap and his group. 

Bridgman, however, formulated the criterion of meaning in a 
much more concrete way. He did not restrict himself to the general 
prescription of how to investigate the meaning of a statement but 
investigated elaborately the operations that have to be carried out in 
order to define the meaning of physical terms like “length” or “thermal 

44 



historical background 


capacity.* By these investigations he found, for example, that the 
operations by which one can distinguish between heat conduction 
and heat radiation, or between heat supplied and mechanical work 
supplied, cannot be performed in all cases. Only if the phenomena 
investigated are of a simple type are these operations feasible. There- 
fore, terms like “heat conduction” or “mechanical work” do not have 
a meaning under all circumstances. 

Bridgman has contributed much to the new philosophy which has 
developed along with twentieth-century science. He has advocated 
strongly the view that the domain of phenomena within which a word 
has meaning is restricted, and no word has meaning if we do not in- 
dicate the circumstances under which it is used. Bridgman has also 
pointed out repeatedly that this new “semantic” aspect is bound to 
have great repercussions upon discourse in politics and religion.” 

In 1931 Carnap was appointed professor of natural philosophy at 
the University of Prague. I succeeded in bringing about this appoint- 
ment, despite the strong opposition of the adherents of traditional 
philosophy, because of a happy coincidence. At that time the Faculty 
of Arts and Sciences was divided into a Faculty of Humanities and a 
Faculty of Science. All professors of philosophy were in the Faculty 
of Humanities and the Faculty of Science gave no instruction in 
philosophy. The president of the Czechoslovakian Republic, Thomas 
G. Masaryk, who had himself been a professor of philosophy, believed 
strongly in the educational value of philosophy. He insisted that the 
Faculty of Science should have a philosopher of their own. I suggested 
Carnap, and as there was no advocate of traditional philosophy left, 
the science faculty agreed. From 1931 on we had in this way a new 
center of “scientific world conception” at the University of Prague. 


22. The Spreed of Logical Empiriciim 

My own interest, which had been for a long time diverted from the 
problems offered by the philosophy of science, returned now to the 
object of my earlier years. The intellectual situation was now in a 
certain respect a similar one. The new science of quantum theory gave 
“ P. W. Bridgman, Yale Review 34, 444 (1945). 

45 



modern science and its philosophy 


rise to a repetition of the crisis that had been precipitated about 1905 
by the relativity theory, but with even greater intensity. Again it was 
maintained that scientific method had failed. The new theories do not 
even claim to give an “explanation” of the physical phenomena. They 
claim only to offer mathematical formulas from which the observed 
phenomena can be derived. The “explanation” is left as a field for 
metaphysical theories, which would claim to give the “real causes” of 
1 things. The argument went mostly that relativ ity as well a s quantum 
J theory give mathematical patterns without any causal justification. 

Remembering our old arguments in the Vienna coffeehouses 
around 1907 about Abel Rey, Ernst Mach, and Henri Poincar6, I 
devoted some work to applying the newly developed “scientific world 
conception” to overcome the new crisis. I tried to show that there is 
not the slightest reason to see in twentieth-century theory an argu- 
ment for anjdealistic o r spiritualis tic world conception, and that this 
opinion only arises from a lack of scientific formulation of the new 
physical theories. This lack has its source in the poor training of 
physicists in philosophy, which makes them often faithful believers in 
the metaphysical creeds imbibed in their early youth “from a nurse or 
a schoolmaster.” “ In Chapters 6 and 7 of the present book I at- 
tempt to give an analysis of physical theories on die basis of the new 
ideas; the scientific argument is carried through consistently without 


suddenly breaking into vague metaphysical discourse. While these 


chapters are more or less devoted to a 
physics. Chapters 8 and 9 contain a ! 


Iphysics, Chapters 8 and 9 contain a special discussion of modem 


q uantum the ory from the same viewpoint, including some remarks on 
“d^erminism” that take up, in a new way, the problem of Chapter 1. 

'if we look from another angle upon those misinterpretations of 
physical theories, it is evident that they are not the result of some 


intellectual inability. Their real source is the urge to find support for 
a metaphysical creed that, for some reason, one cherishes. And this 


reason is, as we have already hinted, the fitness of this metaphysical 
creed to bolster up some political or religious creed that one believes 


“A. N. Whitehead, The Principle of Relativity (Cambridge: The University 
Press, 1922). , 


46 



historical background 


to be indispensable for the well-being of mankind. ^This sociological 
aspect has been for many years familiar to me from the discussions of 
our old Viennese group and, in particular, from the attention that I 
later paid to the influence of religious and politica l creeds upon 
scientific theories in two specific cases; Duhem’s presentation of the 
taken by the Roman Church against the Copemican theoiy, and 
Lenin’s attacks on Mach’s conception of physics. I have touched upon 
this aspect in Chapter 5 and have devoted all of Chapter 10 to it. The 
Copemican confl ict is treated in Chapter 13, the specific nature of 
the Soviet philosophy (dialectical materialism) and, in particular, its 
relation to positivism and empiricism are described in Chapter 11. 

At the time of the meeting in Prague (1929) the Vienna Circle 
and Reichenbach’s group in Berlin were a small number of dissident 
people hemmed in by the vast ocean of German school philosophy, 
which was more or less a development of Kantian metaphysics. It was 
considered to be a specific "German philosophy’’— namely, “German 
idealism’’— and philosophies of other types were regarded with a 
certain suspicion as something "un-German” and “foreign.” The 
workers for a "scientific world conception” had no hope of finding any 
considerable encouragement in Germany. Neurath described this 
cultural and historical situation in his small book, “The Development 
of the Vienna Circle and the Future of Logical Empiricism.” “ Neurath 
writes that the Kantian influence had been slight in the Universities of 
Vienna and Prague and that their philosophy avoided the "Kantian 
interlude” and passed directly from Leibniz to modern positivism. He 
continues, "The influences of English and French thinkers are frequent 
and things happen in Austria parallel to what happens in Warsaw, 
Cambridge, or Paris, rather than to what takes place in Berlin.” 

Although German science had developed along international lines 
and no serious scientist would have objected to any foreign influence, 
in philosophy things were different. There was a strong tendency to 
overrate “German idealism” and to minimize British, French, and 

"*0. Neurath, Le DSveloppement du Cercle de Vienne et Vaoenir de Vem- 
pMsme logique, French translation by General Vouillemin (Paris: Hermann, 
1935 ). 


47 



modern science and its philosophy 


American philosophical trends. We felt very soon that th6 future of 
the “scientific world conception” was to break through this wall and 
to make contact with “foreign” philosophies. This feeling turned out 
to be entirely correct and^e met friendly interest on the part of 
American, British, and French thinkers. In the United States there 
was a natural common ground, the work of the American pragmatists, 
in particular of C. S. Peirce. Charles W. .Morris had cultivated es- 
pecially the ties between pragmatism and the Central European 
positivism. He coined for the result of the very close cooperation of 
these groups the name “logical empiricism,” which seems to me to 
denote the salient point better than any other name. E. Nagel (now 
at Columbia) and W. V. Quine (now at Harvard) came to Vienna and 
Prague, 'as Morris (now at the University of Chicago) had done, to 
make personal contact with Schlick, Carnap, and the other workers in 
this field. 

^n Great Britain our natural link was L. Wittgenstein, whose own 
education had been half Austrian and half British, and, through him, 
Bertrand Russell. The brilliant young Oxford philosopher, A. J. Ayer, 
also came to Vienn^ and published the most readable book on logical 
empiricism that has been written in English, and perhaps the most 
readable book altogether." 

(in France, the advocates of the “new positivism” were naturally in- 
terested in our movement. L. Rougier started his philosophic work on a 
basis similar to that of Schlick) He took his start from Poincare, tried 
to integrate Einstein into the “new positivism,” and wrote the best all- 
round criticism of the school philosophy that I know of, "The paralo- 
gisms of rationalism.” " 

Marcel Boll, an able physicist himself, saw in the Viennese move- 
ment a valuable contribution to a renewed and more vigorous positiv- 
ism and a support of progressive thinking and acting. He translated 
writings of Carnap, Reichenbach, Schlick, and myself into French. 

We encountered also a strong interest in a group of French neo- 

J. Ayer, Language, Truth, and Logic (London: Gollanc?:; New York: 
Oxford University Press, 1936). 

®®L. Rougier, Les paralogismes du rationcdisme (Paris: Alcan, 1920). 

48 



historical background 


Thomists. great influence of the neo-Thomist, P. Duhem, upon the 
Viennese group repeated itself now in reverse. The French general, 
Vouillemin, recommended our group because we replaced the spelling 
“Science” modestly by “science.” He also translated several papers of 
the Vienna Circle into French and published a small book, “The Logic 
of Science and the Vienna School,” ” in which he gave his interpreta- 
tion of logical empiricism. The French neo-Thomists of this group saw 
in logical positivism the destroyers of idealistic and materialistic meta- 
physics, which they regarded as the most dangerous enemies of 
Thomism. 

To organize this international cooperation a preliminary conference 
was held in 1934 in Prague, at which Charles Morris and L. Rougier 
participated. The ground was laid for arranging international con- 
gresses “for the unity of science,” which were to be held every year. 
They actually met in 1935 in Paris, 1936 in Copenhagen, 1937 in 
Prague, 1938 in Cambridge, England, and 1939 in Cambridge, Massa- 
chusetts, at Harvard University. 


23. Teaching fhe Philosophy of Science of Harvard 

In the year 1936, just while the Congress for the Unity of Science 
was in session at Copenhagen, Professor Schlick was assassinated near 
his lecture hall in the University of Vienna by a student. At the court 
trial the attorney for the defendant pleaded extenuating circumstances 
because the student was indignant about Schlick’s “vicious philosophy.” 
Everyone who knew Schlick had been full of admiration for his noble, 
humane and restrained personality. The political implications of the 
expression “vicious philosophy” were obvious. The student received 
a ten-year prison term. When, however, two years later, the Nazi troops 
occupied Vienna and arrested a great many people, Schlick’s murderer 
was released from prison. 

The shots directed at Schlick were a dramatic indication of the dis- 
persal of the Central European positivism that was taking place under 
the pressure of the advancing Nazi power. At the end of 1938 this 

Le general C. E. Vouillemin, La Logique de la science et rdccle de Vienne 
(Paris: Hermann, 1935). 


49 



modern science and its philosophy 


process was completed. ^By far the greatest part of the Central 
Europeans who had worked along the lines of logical positivism had 
left their countries. The immediate reason was either to escape politi- 
cal persecution, or, in many cases, just the feeling that under the 
dictatorship of the Nazis there would be no place for a philosophy 
guided by logic and experience. The majority of the emigrants have 
lived since in the United States, a smaller part in Great Britain^ 

When I arrived in the United States in October 1938, I started a 
lecture tour during which I spoke at twenty-odd universities and col- 
leges on the philosophic interpretations of modem physics. Chapter 7 
of this book presents one of these lectures. 

Since the fall of 1939 I have had the privilege of teaching at 
Harvard University not only mathematical physics but also the philoso- 
phy of science. This teaching has been a great experience for me 
and has been of great influence on my philosophical writing. I started 
with an audience of about fifteen students. Since this was an unusual 
subject I did not quite know what to tell them. I began by presenting 
to them the logical structure of physical theories as envisaged by 
logical empiricism. But very soon I noticed that this was not the 
right thing to do. The frequent discussions that I had with the stu- 
dents showed me what they really wanted to know. By a process of 
interaction, a program was finally worked out that was a compromise 
between what I wanted to tell the students and what they wanted 
to know. 

At that time Harvard University set up its “General Education 
Program.” This was a great help to me since it was based on the 
deficiencies of the traditional curriculum as felt by the students, the 
faculty, and the general public. President J. B. Conant urged, in par- 
ticular, a new approach to the teaching of science. He stressed that 
science teaching has to be linked up with a presentation of the his- 
torical, cultural and psychological background of the work done by 
the great scientists. This program was later developed in Mr. Conant’s 
book. On Understanding Science.^^ 

“J. B. Conant, On Understanding Science (New Haven: Yale University 
Press, 1947). 


50 



historical background 


Stimukted by all these factors and particularly by the rapidly in- 
creasing interest of the students in the philosophical and cultural 
implications of science (I had later an audience of more than two 
hundred fifty), it became more and more clear to me how to work out 
a program for my students. I now put the greatest emphasis on pre- 
senting physics, and science in general, as part of our general pattern 
of thinking and acting. I presented it on one hand as a logical system 
that has to be checked by physical experiments and on the other hand 
as one of the means of expressing man’s attitude towards the world, 
the small world of society and politics and the large world that is our 
astronomical universe. This more historical approach has been familiar 
to me since my student years from the meetings with my older friends. 

All my papers written after 1940 follow this line. Chapter 11 of 
this book describes the interaction between the advance in science and 
the changes in metaphysical systems. Its point is that a great many 
metaphysical systems are merely abandoned systems of science. In 
Chapter 12 the Copemican conflict is discussed from the same view- 
point, particular stress being laid on the role that could be played by 
the intervention of political powers. Chapters 14 and 15, which are 
closely connected with each other, discuss directly the role that in- 
struction in the philosophy of science is to play in the college cur- 
riculum. By the time I came to write them, I had collected a great deal 
of experiential material about this problem. My point was now that 
the philosophy of science should, on one hand, give to the science 
student a more profound understanding in his own field, and on the 
other hand, be for all students a link between the sciences and the 
hiunanities, thus filling a real gap in our educational system. 

In all my writing before 1947 I had stressed the point that science 
gives no support to metaphysical interpretations, of whatever type. I 
had discussed these interpretations only as reflecting the social en- 
vironment of the philosopher. However, after that time, as a result of 
contact with my students and fellow teachers, I became more and 
more interested in the question of the actual meaning of the meta- 
physical interpretations of science— idealistic, materialistic, relativistic, 
and others. For the fact that a great many scientists and philosophers 


51 



modern science and its philosophy 


advance such interpretations and cherish them is as firmly established, 
by our experience, as any fact of physics. 

I now began a new series of investigations into the meaning of 
metaphysics within the framework of logico-empirical and socio- 
psychological analysis. The first preliminary result of these investiga- 
tions is published in Chapter 16 of the present book. 

In 1940 Professor Harlow Shapley, Director of the Harvard College 
Observatory, introduced me into the Conference of Science, Phi- 
losophy and Religion, which meets e%'ery year. This Conference is a 
group of philosophers, educators, social workers, and ministers of all 
denominations, and includes a sprinkling of scientists. They are in- 
terested particularly in the contributions that each of these fields can 
make to the understanding and supporting of the democratic way of 
life. The leaders in this group have been, besides Dr. Shapley, Dr. Louis 
Finkelstein, President of the Jewish Theological Seminary of America, 
and Dr. Lyman Bryson, Director of the Educational Department of the 
Columbia Broadcasting System. From these meetings 1 have learned 
more about the attitudes of different groups of people toward science, 
and what points in the philosophy of science support or are believed 
to support a certain way of life. The large majority in these meetings 
has been rather critical of the scientific outlook and its contribution 
to human welfare. 

I addressed this group several times between 1940 and 1947. My 
contributions centered mostly around the question of whether the 
“relativism” of modem science is actually harmful to the establishment 
of objective values in human life. I made an argument to prove that 
the “relativism of science” has also penetrated every argument about 
human behavior. “Relativism” is not responsible for any deterioration 
of human conduct. What one calls “relativism” is rather the attempt to 
get rid of empty slogans and to formulate the goals of human life 
sincerely and unambiguously. My contributions to these meetings will 
be published “ in due time, under the title “Relativity— a richer truth.” 

^ Beacon Press, Boston. 


52 



CHAPTER 


1 


experience and the law of causality 

T he French mathematician Henri Foincar4, in two books ^ on the 
philosophy of Science, “Science and Hypothesis,” and “The Value 
of Science,” presented the point of view that many of the most 
general principles of theoretical science— such as the law of inertia, 
and the principle of conservation of energy—about which one often 
wonders whether they are of empirical or a priori origin, are actually 
neither of these, but purely conventional definitions depending on 
human arbitrariness. 

The purpose of the present paper is to extend this conception to 
the principle that is in a certain sense the most general in all theoretical 
science, the law of causality. The direct stimulus for this undertaking 
was given by a book which, as a matter of fact, follows an opposite 
tendency, the sagacious and in many respects unprejudiced work, 
“Concepts and Principles of Natural Science,” by Hans Driesch.® The 
author sets out to show that the principle of conservation of energy 
has an a priori nucleus which is no other than the law of causality in 
its precise formulation. In order to demonstrate the apriorily of the en- 
ergy principle, Driesch presents a series of ingenious arguments which 

Foincare, Science and Hypothesis (Science Press, New York, 1905; 
original, 1902); The Value of Science (Science Press, New York, 1907; original, 
1905). 

Driesch, Naturbegr^e und NatururteUe (Leipzig, 1904). 

53 



modern science and its philosophy 


show that experience can never disprove the principle in' question. 
This array of arguments calls to mind in an astonishing way that used 
by Poincare for his conception of the energy principle as a convention. 
The agreement is all the more striking since evidently neither author 
was influenced by the other. Although the conclusions reached are 
quite different, the two arguments are very similar. Since that of 
Driesch is to a large extent applicable to the law of causality, I have 
found in it new support for my conception of this law, which I felt to 
be a necessary consequence of the arguments presented in Poincare’s 
papers. 

The thesis that we shall try to prove states that the law of causality, 
the foundation of every theoretical science, can be neither confirmed 
nor disproved by experience; not, however, because it is a truth known 
a priori, but because it is a purely conventional definition. VVe shall take 
as a basis the form of the law of causality that is freest from undefined 
and ambiguous expressions and contains only the essentials, referring 
directly to the data of the senses. 

The law states that if, in the course of time, a state A of the universe 
is once followed by a state B, then whenever A occurs B will follow it. 

This statement contains everything that is the real content of the 
law of causality. It is important to understand that the law can be 
applied only to the whole universe and not to a part of it. This, how- 
ever, makes it impossible to test the law empirically. In tlie first place, 
one can never know the state of the whole universe, and in the second 
place, it is in general not certain whether it is possible for a state A 
of the universe ever to return. If no state A could ever be repeated, the 
law would be meaningless theoretically, since it refers only to recurring 
states. 

Fortunately, it is not the exact law of causality itself which finds 
application in science, but a formulation of it that asserts only some- 
thing approximate. This says that if, in a finite region of space, die 
state A is at one time followed by the state B and at another time by 
the state C, we can make the region sufficiently large by adding to it its 
environment that the state C becomes as close to the state B as we 
please. 


54 



experience and the law of causality 


In otKer words, in finite systems the law of causality is the more ^ 
nearly valid the larger the system. In the application of the law to a 
finite system, the answer to the question whether the system is large 
enough depends on the degree of accuracy required for the occur- 
rence of the predicted effect. This can be shown by a simple example 
taken from astronomy, the science that has been worked out most 
rationally. Let us consider the system consisting of the sun and the 
earth. A state A of this system, defined by a particular distance of 
separation and relative velocity of the two bodies, is always followed 
by the same series of states, no matter how often A is repeated; but we 
must not take the word “same” too exactly. For, in reality, the series of 
states following A also depends on the distances and velocities of all 
the other planets and the fixed stars as well, and we must include them 
in the system. The more celestial bodies we include, the more ac- 
curately is the law of causality obeyed. However, if we take into the 
system only the planets with their satellites, the accuracy is sufficient 
for all practical purposes. We see from this example that there do exist 
finite systems to which the law of causality is applicable. 

Whether a given system will behave in this way cannot be known 
beforehand; for this reason the so-called inductive method has been 
developed. When an investigator sees “that in a system the state B 
follows the state A, not once but often, he will say that A is the cause 
of B.” This means nothing else, however, than that the law of causality 
is applicable to the system under discussion. For an all-embracing 
system, a single case in which B follows A is enough to enable one to 
draw conclusions for all times thereafter. For a finite system, however, 
it is necessary to decide in every case whether the law of causality is 
applicable. Naturally, such a decision can never be a final one, for the 
law of causality as applied to finite systems is not the real one, but only 
a substitute for it. The real law itself is only an ideal which the law for 
finite systems approaches as a hmit as these systems are made larger 
and larger. ' 

Here we do not wish to concern ourselves with the difficulties 
arising from the finiteness of empirical systems. It will soon appear that 
these are relatively unimportant in comparison with arguments that 


55 



modern science and its philosophy 


would place the law of causality in a peculiar light even if it were 
rigorously valid for finite systems, which will be assumed hereafter. 
We have, then, a finite system for which the law is valid that if the 
state A is followed once by the state B, then A will be followed every 
time by B. In this statement there occurs, however, a single word— 
“state”— that is not explainable directly through any reference to sense 
data. And the analysis of this word will suffice to demolish this seem- 
ingly so strongly built meaning of the law. 

What is a “state of a system?” The most obvious explanation would 
be that by state we mean the ensemble of the perceptible properties of 
a system. This would be a clear meaning. However, if we take the 
word "state” in this sense, the law of causality becomes incorrect, as 
can be seen from simple examples. Let us assume that the system con- 
sists of two iron rods, lying side by side on the table, that is, in state A. 
Left to themselves, they will continue to lie quietly; that is, A is fol- 
lowed by A. If now we replace one of the iron rods by a magnetized 
one of exactly the same appearance, the initial state, according to our 
definition of the word “state,” will be the same as before, namely, A. 
The rods will now move toward each other, however; that is, A is now 
followed, not by A, but by B. In order to be able to say that the law of 
causality is still valid, we must say that the initial states were only ap- 
parently the same. We must include in “state” not only the totality of 
perceptible properties, but also another, namely, in our example, 
magnetization. A property belonging to the definition of the state is 
called a state variable of the system. 

How does it come about that we assign to bodies imperceptible as 
well as perceptible properties? Such properties— electric charge, chemi- 
cal affinity, and so on— are characteristics that indicate how the body 
possessing them behaves when brought into certain situations. They 
are, according to Driesch, "the aggregates of possibilities, regarded 
as reality.” 

This, however, means only that if a body in a given situation be- 
haves differently from another body the state of which, in the sense 
first defined, is the same, we assign to it new state variables in addition 
to the perceptible ones. This in turn means only that if the law of 


56 



experience and the law of causality 


causality is not valid according to one definition of the state, we re- 
define the state in such a way that the law is valid. If that is the case, 
however, the law, which appeared to be stating a fact, is transformed 
into a mere definition of the word “state." We can express the law in 
the following form; “By the word ‘state’ one understands the per- 
ceptible properties of a system of bodies, plus a series of fictitious 
properties, of which so many are included that the same states are 
always followed by the same states.” In this form the “law of causality” 
no longer looks at all like a law. 

Because of the fact that the word “state,” which occurred in that 
form of the law of causality first taken as a basis, is only defined by 
this law, the latter loses the character of a factual proposition and 
becomes a definition. Of a definition, however, one cannot say that it 
is empirical or a priori; it is only a product of human imagination. , 

The conclusion to which the foregoing refiections lead is that “the 
law of causality is only the establishment of a terminology.” Because 
this law forms the basis of the whole of theoretical science, the latter 
itself is also nothing else than a suitably chosen terminology. Whereas 
experimental science describes the properties of bodies as given by our 
senses, and the changes in these properties, the task of theoretical 
science is to provide bodies with fictitious properties the chief purpose 
of which is to insure the validity of the law of causality. Theoretical 
science is not research but a sort of remodeling of nature; it is work of 
the imagination. From this it is clear whence the so-called pure— that is, 
a priori— science, the possibility of which led Kant to write his Critique 
of Pure Reason, derives its conviction of being right. The principles of 
pure science, of which the foremost is the law of causality, are certain 
because they are only disguised definitions. 

Pure science states nothing about empirical nature; it only gives 
directions for portraying nature. All the arguments which Driesch has 
arrayed so ingeniously for the existence of a piure science show indeed 
that there are principles independent of our experience, but fail to 
explain why this is the case. The reasons are completely revealed by 
the conception presented above. 

i^Tbus we see that the latest philosophy of nature revives in a striking 

57 



modern science and its philosophy 


way the basic idea of critical idealism, that experience only serves to 
fill in a framework which man brings along with him as a part of his 
nature. The difference is that the old philosophers considered this 
framework a necessary outgrowth of human organization, whereas we 
see in it a free creation of human imagination. 

How often the question has been put: “How can it happen that 
man can work out all of outer nature, which is, after all, completely 
independent of his mind?” Are not nature and the human intellect in- 
commensurable things? From our standpoint it is easy to answer that 
the nature which the human mind rationalizes by means of theoretical 
science is not at aU the nature that we know through our senses. The 
law of causality and with it all of theoretical science have as their 
object not empirical nature but the fictitious nature of which we spoke 
above. The latter, however, is not only the object, but also the work 
(and work, not in any metaphysical sense, but in the ordinary sense) 
of man; hence it can of course be completely comprehended by him. 

To the fundamental questions of theoretical science, experience and 
experiment can never give an unequivocal answer. If I wish, I can 
provide all bodies with state variables that are all qualitatively dif- 
ferent, in order to fulfill the law of causality. I can regard heat, elec- 
tricity, magnetism, as properties of bodies, essentially different from 
one another, just as is done in modem energetics, and as Driesch does. 
On the other hand, if I wish, I can get along without introducing quali- 
tative differences. For example, I can introduce only the motion of 
masses; but then, in order to obtain the necessary diversity, I must take 
refuge in unconfirmable hidden motions. This leads to the purely me- 
chanical picture of the world, which Democritus dimly conceived as 
an ideal, and which occurs mostly in the form of atomism. This purely 
quantitative picture of the universe, striving to manage with a mini- 
mum number of qualities, was given its most logical development in 
the book ‘Thilosophy of Inanimate Matter” by Adolf Stohr,® where 
even the qualitative specificities still adhering to mechanical atomism 
were suppressed in favor of purely geometrical-quantitative schemes. 
This work, as the most radical carrying out of the program of the 
* A. St&hr, Phttosophie der unbelebten Materie (Leipzig, 1907). 

58 



^ experience and the law of causality 

atomists, occupies a significant place in the literature of natural 
philosophy. In a quite different way, H. A, Lorentz and his pupils have 
created a quantitative world picture by breaking away from the mech- 
anistic tradition and introducing as state variables electric charge 
and electric and magnetic field intensities. Thus arose the electro- 
magnetic picture of the world. Among all these, it is not possible to 
choose uniquely on the basis of experience. One may be simpler, an- 
other more complicated, but none true or false. 

We see that it is not at all a scientific question, in the narrower 
sense, what world picture I make for myself. The world pictures are 
only more or less different expressions used for the same thing— 
empirical nature. 

Furthermore, the same is true of a question which has long held 
sway as a so-called question of world conception, the answer to which 
is above all demanded of the scientist. It is a question that seems to 
carry with it an i nfi nite number of emotional values, and yet it is only 
a question of terminology. It is the question (to return at the end to 
the book of Driesch mentioned at the beginning) whether the phe- 
nomena of animal and plant life can be explained by means of the laws 
of physics and chemistry, the question that is usually summarized in 
the high-sounding expression, “vitalism or mechanism?” 

We are indebted to Driesch for the first clear and unprejudiced 
formulation of the problem which, following him, and with reference 
to what has been said hitherto, we can express as follows: Must we, in 
order to satisfy the law of causality in the domain of life, ascribe to 
the body besides the properties (state variables) of physics and 
chemistry, also other, qualitatively different properties? Driesch tries 
to show that we must do this, and introduces entelechy, as a state 
variable peculiar to living bodies. This attempt of Driesch’s to show 
that it is impossible to get along with the state variables of physics and 
chemistry alone seems to me not entirely convincing. To be sure, 
Driesch shows that we can assume for the living processes a specific 
state variable, but not that we must. For it is not possible to foresee 
every trick that one might invent in the fiction of hidden combinations 
of inorganic state variables. In favor of vitalism^ I should like to remark 


59 



modern science and its philosophy 


that, just as I cannot force someone who regards heat as a specific state 
variable to consider it as a motion of mass particles, so 1 cannot force 
the adherents of entelechy to replace it by fictitious state variables. 
However, that is not very important for the purpose of the present 
work. What is important is that, from the bio-theoretical works of 
Driesch, if we examine them from the standpoint adopted by us, it is 
clear that the question “vitalism or mechanism?” is not a question of 
fact. It is not a question to which a crucial experiment can answer yes 
or no. It is rather a question the solution of which depends on the in- 
genuity of the human imagination and can never be convincing to all 
men. The question is not: “Is that thus or so?” It is rather: “Can we 
paint the picture in this or in that style, or in both?” 

With the question of world conception in the ethical-religious sense, 
all this has nothing whatsoever to do. 


60 



CHAPTER 



the importance for our times of Ernst Mach’s 
philosophy of science 

T hebe is something remarkable about the teachings of Mach. 
Philosophers often ridicule or disdainfully reject them as the 
work of a physicist dabbling in philosophy; physicists often 
deplore them as aberrations from the right path of respectable, realistic 
natural science. Yet neither physicists and philosophers, nor historians 
and sociologists, nor many others, can get rid of Mach. Some attack 
him passionately; others extol him with fervor. There is something 
fascinating about his simple, straightforward teachings. In spite of 
their simplicity they are stimulating and provocative. There are in- 
deed but few thinkers who can provoke such sharp differences of 
opinion, who are so inspiring to some and so utterly repugnant to 
others. What is there in these doctrines that makes it impossible for 
anyone, whatever his views may be, to avoid adopting some definite 
attitude towards them? 

This is what I should like to discuss in the present paper. I have 
formed a fairly definite opinion about the position that Mach occupies 
in the intellectual life of our times, and this position, I believe, will ex- 
plain why the battle rages about him so furiously. It is not a question 
here of the details of Mach’s teaching, often individually and his- 
torically conditioned, but rather of their nucleus, which is just the 
focus of the struggle. I will not speak therefore about the general at- 


61 



modern science and its philosophy 


titude of Mach towards the psychophysical problem, nor about his 
separate contributions to physics and psychology, but only about his 
conception of the tasks and possible aims of exact science. 

In recent years, among creatively active physicists and mathema- 
ticians, there has become noticeable a reaction against the conceptions 
of Mach. When one of the most outstanding theoretical physicists of 
our time. Max Planck,^ and one of the foremost living geometricians, 
E. Study," characterize these conceptions as being partly misleading, 
partly incapable of being carried out, and partly actually harmful for 
science, the fact provides food for thought and cannot be lightly 
brushed aside. 

What an investigator with the markedly constructive talents of 
Planck dislikes above all in the views of Mach is his judgment of values. 
•. For the investigator, every new theory that is supported by experiment 
is a piece of newly discovered reality. According to Mach, on the other 
hand, physics is nothing but a collection of statements about the con- 
nections among sense perceptions, and theories are nothing but eco- 
nomical means of expression for summarizing these connections. He 
says: 

/ 

' The aim of natural science is to obtain connections among phenomena. 
Theories, however, are like withered leaves, which drop off after having 
enabled the organism of science to breathe for a time.” 

This phenomenalistic conception, as it is called, was familiar to 
Goethe. In his posthumous “Maxims and Reflections” he says: 

Hypotheses are the scaffolds which are erected in front of a building and 
removed when the building is completed. They are indispensable to the 
worker; but he must not mistake the scaffolding for the building. 

And still more drastically: 

^ M. Planck, Die Einheit des physikalischen Weltbildes (Leipzig, 1909). 

”E. Study, Die realistische Weltansicht und die Lehre vom Raume (Bruns- 
wick, 1914). 

” E. Mach, Die Qeschichte und die Wurzel des Satzes von der Erhaltung der 
Arbeit, written in 1871. An English translation, History and Root of the Principle 
of the Conservation of Energy, was published in 1911. 


62 



Ernst Mach's philosophy of science 


constancy of phenomena alone is important; what we think about 
them is quite immaterial. 

It will be said, however, that Goethe was not really a good physicist, 
and that we can see in his case an example of how such basic prin- 
ciples hinder the spirit of research. Thus Planck says; 

When the great masters of exact investigation of nature gave their ideas 
to science, when Nicholas Copernicus removed the earth from the center 
of the universe, when Johannes Kepler formulated the laws named after 
him, when Isaac Newton discovered gravitation . . . —the series could be 
long continued— surely, economical points of view were the very last thing 
to steel these men in their struggle against traditional opinions and dominat- 
ing authorities. No, it was their unshakable belief— whether resting on an 
artistic, or on a religious basis— in the reality of their world picture. In view 
of these certainly incontestable facts, one cannot reject the surmise that, if 
the Mach principle of economy were really to be put at the center of the 
theory of knowledge, the trains of thought of such leading spirits would be 
disturbed, the flight of their imagination crippled, and consequently the 
progress of science perhaps fatefully hindered.* 

That these fears in such generality are groundless can be readily 
seen if one recalls the views of one of the greatest theoretical physicists 
of the nineteenth century, James Clerk Maxwell, on the nature of 
physical theories. One need only read the introduction to his essay on 
Faraday’s lines of force (1855) ® to be convinced that he was com- 
pletely an adherent of the phenomenahstic standpoint. Yet one cannot 
say of him that such adherence in any way crippled the flight of his 
imagination; indeed, quite the opposite. The conception of the rela- 
tive worthlessness of the theory in comparison to the phenomenon 
gives to the theorizing of such an investigator something especially free 
and imaginative. 

I am willing to concede that the phenomenalistic doctrine is 
pleasing to those whose work in physics is descriptive rather than con- 
structive. Many such people, who are capable of describing very neady 

* Planck, op. cit., p. 36. 

° Published in Cambridge Philosophical Society Transactions, 1864, and in 
The Scientific Papers of James Clerk Maxwell, ed. by W. D. Niven (Cambridge: 
The University Press, 1890), vol. 1. 


63 



modern science and its philosophy 


definite— even if very special— phenomena, may regard themselves be- 
cause of this doctrine as being sublimely superior to the imaginative, 
creative spirit, whose works, after all, are only phantoms and 
“withered leaves.” I do not believe, however, that for these people the 
philosophy of Mach has crippled the imagination. Rather, it is the case 
of an imagination crippled by nature, which uses the teachings of Mach 
as a beautiful cloak to cover its deformity. It may have been cases like 
these which cause Planck, at the end of the lecture cited above, to 
hurl at the preachers of the phenomenalistic doctrines the Biblical 
words: “By Aeir fruits ye shall know them.” 

Concerning this criterion of the fruits I shall have more to say. 
First, in connection with the same Biblical allusion, I should like to 
introduce a quotation from Pierre Duhem about the value of physical 
theories. Duhem was the most outstanding representative in France 
of ideas similar to those of Mach. He said: 

By the fruit one judges the tree; the tree of science grows exceedingly 
.slowly; centuries elapse before one can pluck the ripe fruits; even today it is 
hardly possible for us to shell and appraise the kernel of the teachings that 
blossomed in the seventeenth century. He who sows cannot therefore judge 
the worth of the corn. He must have faith in the fruitfulness of the seed in 
order that he may follow untiringly his chosen furrow when he casts his ideas 
to the four winds of heaven.” 

These remarks of the greatest and most accurate student of the his- 
tory of physics are perhaps the best answer to the opinion expressed 
by Planck 

that even our present world picture, although it shows the most varied colors 
according to the individuality of the investigator, nevertheless possesses 
certain features that can never be obliterated by any revolution, either in 
nature or in the human mind.^ 

These enduring features arise, according to Mach, from the fact 
that all possible theories must give the same connection between phe- 

•P. Duhem, L’&oolution de la micanique (Paris, 1903), translated into Ger- 
man by Ph. Frank and E. Stiasny as Die Wandlungen der Mechanik (Leipzig, 
1912). 

^ Op. ctt., p. 35. 


64 



Ernst Mach's philosophy of science 


nomena; this very fact guarantees a certain constancy. The known 
connections among phenomena form a network; the theory seeks to 
pass a continuous surface through the knots and threads of the net. 
Naturally, the smaller the meshes, the more closely is the surface fixed 
by the net. Hence, as our experience progresses the surface is per- 
mitted less and less play, without ever being unequivocally determined 
by the net. 

Since, according to Planck and Study, the basic principles of Mach 
would do nothing but harm to physics, it is fortunate for physics that 
these principles have never been thoroughly applied by their adherents, 
even if it is a gloomy sign for the principles themselves. Thus Study 
says about positivism, as he calls the doctrines of Mach: 

We regard this principle as a perfect utopianism. The possibility of its 
existence is based entirely on the fact that it is disavowed by its own fol- 
lowers at every step. Up to now there has never been any serious attempt 
to apply it consistently . . . We are dealing here with a question of prin- 
ciple and must therefore distinguish between the theory of positivism and 
the practice of the (fortunately for themselves) thoroughly inconsistent 
positivists.® 

Planck says similarly: 

We then attain a more realistic mode of expression . . . which is actually 
the one always used by physicists when they speak in the language of 
their science.® 

With biting sarcasm. Study says; 

In numerous cases the hypotheses that are basely denounced at the 
official reception (why not atomistics, too?) are admitted, under a different 
name and through a back door especially arranged for this, into the sanctuary 
of science. Such names and corresponding motivations are by no means few. 
Without any effort the writer collected a full dozen of them: “most com- 
plete and simplest description” (Kirchhoff) . . . “subjective means of re- 
search,” “requirement of conceivabiKty of facts,” “restriction of possibilities,” 
“restriction of expectation,” “result of analytic investigation,” “economy of 
thought,” “biological advantage” (all of these employed by E. Mach).“ 

® Study, op. cU., pp. 36, 41. 

® Quoted by Study, p. 37. 

“Ibid., p. 37. 


65 



modern science and its philosophy 


With equal mockery, Planck remarks: 

I should not be at all surprised if a member of the Mach school were to 
come out some day with the great discovery, that . . . the reality of atoms 
is just what is required by scientific economy 

Other authors, too, point out the glaring contradiction that exists 
among the admirers of Mach between theory and practice. A peculiar 
theory of the nature of physical theories is set up, but as soon as physics 
really begins the positivist behaves practically like any other physicist. 
A follower of Mach can proclaim that physics has to do only with re- 
lations among sense perceptions, but the preacher of this doctrine 
speaks as a physicist exactly like any one else, about matter and energy, 
and even about atoms and electrons. 

However, it is just this apparently so palpable contradiction that 
can lead to the understanding of the permanent nucleus of Mach’s 
teachings. Let us listen once more to Study: 

The whole situation is a striking reminder of Kronecker’s proposal to 
abohsh the irrational numbers and to reduce mathematics to statements 
about integers; in this case, too, the suggestion has remained programmatic, 
and for the same good reasons.*^ 

The analogy, as I see it, is a very appropriate one. But I should like to 
give it a different interpretation from that of Study. It is obviously 
pointless actually to express all theorems of mathematics as theorems 
about integers. In principle, however, it is highly enlightening to know 
that all theorems about irrational numbers, and hence also all theorems 
about limiting values, could be expressed as theorems about integers. 
Once this possibility has been substantiated, the whole of analysis can 
proceed to develop as usual. But now when a theorem about deriva- 
tives is set up and somebody begins to subtilize about it, asking 
whether this theorem is really in agreement with the “nature” of the 
differential and going into profound and skeptical deliberations con- 
cerning this “nature,” he can be told quite simply: “I could express this 

“ M. Flanck, “Zur Machschen Theorie der physikalischen Erkenntnis," Vtertel- 
jahTsschrift fiir wisaenschaftliche Philosophie und Soziologie 34, 497 (1911). 

“ Study, op. dt., p. 39, 


66 



Ernst Mach's philosophy of science 


theorem, if I took enough time, as a theorem about integers; the nature 
of this theorem is hence no more and no less mysterious than that of 
the natural numbers.” 

The situation is quite similar with respect to Mach’s physical theory 
of knowledge. It is not a question of actually expressing all physical 
statements as statements about relations among sense perceptions. It 
is important, however, to establish the principle that only those state- 
ments have a real meaning that could in principle be expressed as state- 
ments about the relations among our perceptions. To express the law 
of conservation of energy or the law of equipartition of energy among 
all the degrees of freedom as statements about relations among per- 
ceptions is just as laborious, but also just as superfluous, as to express 
the theorem that the derivative of the sine is the cosine as a statement 
about integers. In principle, however, both are certainly possible. 

For the inner working of physics it is in most cases practically im- 
material whether one has adopted Mach’s standpoint or not. Similarly, 
in Kronecker’s lectures on integral calculus there is nothing to be found 
that differs essentially from the presentation of other mathematicians. 

Wherein, then, lies the value of the doctrines of Mach for physics? 

My view is that their main value is not that they help the physicist 
to go forward in his physical work, but rather that they provide the 
means for defending the edifice of physics against attacks from outside. 

One who examines dispassionately the concepts that today are at 
the basis of the system of hypotheses of physics will hardly be able to 
assert seriously that the atom, the electron, and the quantum form 
really satisfactory ultimate building blocks. Every thinker who is some- 
what inclined toward logical thoroughness can find many obscurities 
in these concepts. Into these nebulosities boring doubt can penetrate 
and try to shake the whole system of physics as the foundation of our 
scientific world picture. Here Mach steps in and says: 

All these concepts are only auxiliary concepts. What is significant is ffie 
connection among phenomena. Atoms, electrons, and quanta are only links 
to represent a connected system of science; they make it possible to derive 
logically tlie immeasurable system of connected phenomena from a few 
abstract principles. But these abstract principles are then only the means 


67 



modern science and its philosophy 


to an economical representation. They are not the epistemological basis. The 
reality of physics can never be shaken by any criticism of the auxiliary 
concepts. 

The work of Mach is therefore not essentially destructive, as it is often 
represented to be; positivism is not, as Study calls it, a “negativism,” 
but on the contrary is an attempt to create an unassailable position for 
physics. As a matter of fact, Planck, too, recognizes this when he says: 

To it [the positivism of Mach] belongs in full measure the credit for 
having found again, in the face of tlireatening skepticism, the only legitimate 
starting point of all investigation of nature, the sense perceptions.'^ 

That Planck condemns so sharply the conception of Mach seems to 
me therefore to arise from the fact that he considers it only from the 
viewpoint of its application within physics. It must be said, however, 
that, even when regarded from this standpoint, the phenomenalistic 
conception has already accomplished something and is perhaps capable 
of accomplishing still more. In the boundary regions of physics, where 
general concepts like space, time, and motion play a role, the epistemo- 
logical position one takes is no longer entirely immaterial. Indeed, it is 
universally known today that Einstein’s general theory of relativity 
and gravitation grew immediately out of the positivistic doctrine of 
space and motion, as Einstein himself has discussed in detail in his 
reference to Mach.“ 

On the whole, however, I will gladly concede to Planck and Study 
that positivism itself has not done much in clearing up individual ques- 
tions of physics. But from this it does not follow that positivism is, in 
general, of no value. The "fruits” of Mach’s teachings are indeed not 
purely physical in character. It will be recalled that in recent years an 
attempt was made to take advantage of the criticism of the basic physi- 
cal concepts in order to proclaim the bankruptcy of the scientific world 
picture. Bearing this in mind, one must ascribe a high value to the effort 
of Mach to make physics independent of every metaphysical opinion. 

H. Poincar^ says: 

“ Planck, Die Einheit des phtjsikaliechen Weltbildes, p. 34. 

“A. Eimtein, Physikailische Zeitschrift 17, 101 (1916). 

68 



Ernst Mach's philosophy of science 


At first glance it appears to us that theories last only a day and that 
mins heap up on ruins ... If we examine the matter more closely, how- 
ever, we find that what decays is those theories which claim to teach us 
what things are. But there is something in them which endures. If one of 
them has revealed to us a tme relation, this relation has been acquired for 
all time. We shall find it again under a new cloak in the other theories 
which will reign successively in its place.'® 

In a very determined manner, the French philosopher Abel Rey 
emphasizes the importance for the general intellectual life of preserving 
the edifice of physical ideas. He says: 

If those sciences which, historically, have had an essentially emancipating 
effect go down in a crisis that leaves them only with the significance of tech- 
nically useful collections but robs them of every value in connection with 
the cognition of nature, this must bring about a complete revolution in the 
logical art and in the history of ideas . . . The emancipation of the mind, 
as it has been conceived since Descartes, thanks to physics, is a most fatally 
erroneous idea. One must introduce another way, and restore to a subjective 
intuition, to a mystical sense of reality— in short, to the mystery — everything 
that we believed had been taken away from it . . . If, on the contrary, it 
turns out that there is no justification for regarding this crisis as necessary 
and incurable . . . then the rational and positive method remains the best 
nurse of the human spirit.'® 

Here we have a very clear exposition of the dangers that would 
arise for the whole world conception from a physics that had no other 
epistemologic foundations than those auxiliary concepts which are so 
much exposed to criticism. 

Mach himself saw the real value of his theories in the fact that they 
permitted setting up a connection, as free from contradictions as pos- 
sible, between physics on the one hand and physiology and psychology 
on the other. He who still doubts this has only to read the general sec- 
tions of the Analysis of Sensations.^'' Here it is stressed again and again 
that one must exert oneself to develop physics while using concepts 

'® H. Poincare, La Valeur de la science (Paris, 1905); English translation by 
G. B. Halsted, The Value of Science (New York, 1907). 

'* A. Rey, La ThSorie de la physique chez les physiciens contemporains (Paris, 
1907), p. 19. 

" The Analysis of Sensations and the Relation of the Physical to the Psychical, 
translated by C. M. Williams and Sydney Waterlow (Dhicago, 1914). 


69 



modern science and its philosophy 

that will not have to be given up in a transition to a neighboring branch 
of science. 

From this effort of Mach to employ only concepts that will not lose 
their usefulness outside of physics we can understand his opposition to 
atomistics, which in particular has caused many physicists to turn 
against him, It is true that atomistics, when applied to physiologic and 
psychologic problems, easily leads into a blind alley. Such questions 
arise as: “How can a brain atom think?” “How can an atom perceive 
green, since, after all, it is itself only a miniature picture of a macro- 
scopic body composed of perceptions?” 

I will not deny that Mach allowed himself to be misled by this 
argument into attacking the use of atomistics in physics more sharply 
than can be justified. After all, the usefulness of the atomic theories in 
this limited realm is certainly indisputable. His followers, as is gen- 
erally the case, often saw in this weakness of the master his greatest 
strength, and wished to banish the atom entirely from physics. I believe 
that one can completely free the nucleus of Mach s teachings from this 
historically and individually conditioned aversion to atomistics. The 
■ atoms are auxiliary concepts just like others that can be employed 
advantageously in a limited domain. They are not suitable for an 
epistemological foundation. Once we have adopted this point of view, 
we are all the freer in employing the concept of atoms wherever it is 
admissible. I believe that even Planck would not object so much to the 
kernel shelled in this way. It is then no longer so very strange when 
one declares the atoms, if not their reality, to be a requirement of 
economy. They can be the simplest means of representing physical 
laws without thereby being suitable to form an epistemological founda- 
tion. 

In general, phenomenalism will neither particularly further nor 
hinder the physicist in his field of work. Thus Maxwell, who doubtless 
thought positivistically, wrote the work that laid the foundation for 
the molecular theory of gases. The phenomenalistic conception be- 
comes a danger only in those cases in which the requirement of 
economy is not realized with equal intensity. The most noteworthy 
historical example is perhaps Goethe’s doctrine of colors. However, if 


70 



Ernst Mach's philosophy of science 


one wishes to pass judgment on a person of such strong individuality 
one must not forget, as A. Stohr quite correctly points out, that the 
requirement of economy may mean something quite different for 
every individual.^ For one, it signifies a minimum of hypotheses; for 
another, say, a minimum of different kinds of energy. The former is 
the case for the extreme phenomenalist Goethe; the latter for the pure 
mechanist. 

It will perhaps be instructive, as a comparative case, to recall a 
theoretical physicist who, as an immediate student of Mach, really tried 
to construct a system of physics and chemistry in which no hypothetical 
corpuscles, whether atoms or electrons, occur, and which embraces all 
the phenomena known at present. It cannot be denied that Gustav 
Jaumann in his numerous works has undertaken this task with great 
constructive force.^“ I do not believe, however, that the result has 
turned out to be really in the spirit of Mach’s teachings. To be sure, it 
corresponds to the surface requirement that all atomistics be omitted, 
but it hardly corresponds to the requirement of economy. A large num- 
ber of constants are employed about which the theory can make no 
prediction. The Jaumann system makes it possible only in a very 
limited degree to derive phenomena, also with regard to numerical 
values, from a small number of hypotheses. To show the independence 
of physical research and the epistemologic basis, one can mention that 
the most energetic attempt to refute the corpuscular theory of elec- 
tricity, that of F. Ehrenhaft, has no connection whatsoever with 
philosophic dogmas of any kind. 

I believe that I have now made clear to some extent the significance 
of Mach. In order fully to survey his position in the intellectual life of 
our times, however, we must find a more remote standpoint, in order 
to obtain a better view. 

If we read the most important work of Mach, his Mechanics,^ we 

A. Stohr, PhUosophie der unbelebten Materie (Leipzig, 1907), pp. 16 ff. 

“ G. Jaumann, "Geschlossenes System physikalischer und chemischer DiSeien- 
tialgesetze,” Sltzungsberichte der Wiener Akademie der Wissenschaften, math.- 
scientific class, section Ila (1911), and many other papers in the same journal. 

E. Mach, Die Mechanik in Hirer Entwickelung, historisch-kritisch dargestellt 
(Leipzig, 1883); The Science of Mechanics; a Criticfl and Historical Exposition 


71 



modern science and its philosophy 

that will not have to be given up in a transition to a neighboring branch 
of science. 

From this efiEort of Mach to employ only concepts that will not lose 
their usefulness outside of physics we can understand his opposition to 
atomistics, which in particular has caused many physicists to turn 
against him, It is true that atomistics, when applied to physiologic and 
psychologic problems, easily leads into a blind alley. Such questions 
arise as: “How can a brain atom think?” “How can an atom perceive 
green, since, after all, it is itself only a miniature picture of a macro- 
scopic body composed of perceptions?” 

I will not deny that Mach allowed himself to be misled by this 
argument into attacking the use of atomistics in physics more sharply 
than can be justified. After all, the usefulness of the atomic theories in 
this limited realm is certainly indisputable. His followers, as is gen- 
erally the case, often saw in this weakness of the master his greatest 
strength, and wished to banish the atom entirely from physics. I believe 
that one can completely free the nucleus of Mach’s teachings from this 
historically and individually conditioned aversion to atomistics. The 
atoms are auxiliary concepts just like others that can be employed 
advantageously in a limited domain. They are not suitable for an 
epistemological foundation. Once we have adopted this point of view, 
we are all the freer in employing the concept of atoms wherever it is 
admissible. I believe that even Planck would not object so much to the 
kernel shelled in this way. It is then no longer so very strange when 
one declares the atoms, if not their reality, to be a requirement of 
economy. They can be the simplest means of representing physical 
laws without thereby being suitable to form an epistemological founda- 
tion. 

In general, phenomenalism will neither particularly further nor 
hinder the physicist in his field of work. Thus Maxwell, who doubtless 
thought positivistically, wrote the work that laid the foundation for 
the molecular theory of gases. The phenomenalistic conception be- 
comes a danger only in those cases in which the requirement of 
economy is not realized with equal intensity. The most noteworthy 
historical example is perhaps Goethe’s doctrine of colors. However, if 


70 



Ernst Mach's philosophy of science 


one wishes to pass judgment on a person of such strong individuality 
one must not forget, as A. Stohr quite correetly points out, that the 
requirement of economy may mean something quite different for 
every individual.” For one, it signifies a minimum of hypotheses; for 
another, say, a minimum of different kinds of energy. The former is 
the case for the extreme phenomenalist Goethe; the latter for the pure 
mechanist. 

It will perhaps be instructive, as a comparative case, to recall a 
theoretical physicist who, as an immediate student of Mach, really tried 
to construct a system of physics and chemistry in which no hypothetical 
corpuscles, whether atoms or electrons, occur, and which embraces all 
the phenomena known at present. It cannot be denied that Gustav 
Jaumann in his numerous works has undertaken this task with great 
constructive force.^“ I do not believe, however, that the result has 
turned out to be really in the spirit of Mach’s teachings. To be sure, it 
corresponds to the surface requirement that all atomistics be omitted, 
but it hardly corresponds to the requirement of economy. A large num- 
ber of constants are employed about which the theory can make no 
prediction. The Jaumann system makes it possible only in a very 
limited degree to derive phenomena, also with regard to numerical 
values, from a small number of hypotheses. To show the independence 
of physical research and the epistemologic basis, one can mention that 
the most energetic attempt to refute the corpuscular theory of elec- 
tricity, that of F. Ehrenhaft, has no connection whatsoever with 
philosophic dogmas of any kind. 

I believe that I have now made clear to some extent the significance 
of Mach. In order fully to survey his position in the intellectual life of 
our times, however, we must find a more remote standpoint, in order 
to obtain a better view. 

If we read the most important work of Mach, his Mechanics,^ we 

A. Stohr, PhUosophie der unbelebten Materie (Leipzig, 1907), pp. 16 ff. 

“ G. Jaumann, "Geschlossenes System physikalischer und cliemischer DiSeren- 
tialgesetze,” Sltzungsberichte der Wiener Akademie der Wissenschaften, math.- 
scientific class, section Ila (1911), and many other papers in the same journal. 

“ E. Mach, Die Mechanik in ihrer Entwickelung, historisch-kritisch dargesteUt 
(Leipzig, 1883); The Science of Mechanics; a Critical and Historical Exposition 

71 



modern science and its philosophy 


shall find that in no section does he give us so deep an insight into his 
innermost thoughts and intellectual inclinations as in the wonderful 
chapter on theological, animistic, and mystical points of view in 
mechanics. A wind of refreshing coolness blows from these statements. 
What other authors have treated with vehement blustering, often with 
a quiet announcement of a small auto-da-fe for the opponent, we see 
here discussed in a genuinely scientific spirit. Yet throughout the 
whole chapter there quivers an undertone of suppressed excitement. 
One encounters there that state of being drunk with soberness which 
has been attributed to the Age of Enlightenment. Indeed, Mach descries 
in this age his spiritual home. In the chapter referred to, he says: 

For the first time, in the literature of the eighteenth century enlightenment 
appears to be gaining a broader base. Humanistic, philosophical, historical, 
and natural science come into contact at this time and encourage one an- 
other to freer thought. Everybody who has experienced this upsoaring and 
emancipation, even if only in part, through the literature, will feel a life- 
long melancholy nostalgia for the eighteenth century. 

The personal acquaintances of Mach are aware that he was an 
ardent admirer and reader of Voltaire. One of his former assistants. 
Professor George Pick, informed me that Mach most emphatically con- 
demned Lessing’s attacks on Voltaire. It is known, too, that Josef 
Popper, of whom Mach says that for a long time he was the only man 
with whom he could talk of his physical and epistemologic opinions 
without precipitating a conflict, wrote a whole book devoted to the 
defense, indeed the glorification, of Voltaire. 

In my opinion, Mach was led in this predilection by a correct 
estimate of himself. We can understand the role which as philosopher 
he plays in the intellectual life of the present if we look upon his 
teachings as the philosophy of enlightenment appropriate to our time. 

Since this conception may easily be misunderstood, I must discuss 
it more fully. First of all, the word “enlightenment” has acquired such 
a bad connotation that perhaps many will see in this description a dis- 
paragement of Mach. We must therefore try to make clear something 

of its Principles, translated from the second German edition by Thomas J. McCor- 
mack (Chicago, 1893). 


72 



Ernst Mach's philosophy of science 

about the nature of this enlightenment and the reasons for its sub- 
sequent neglect. 

The first period of enlightenment in modem times began with 
the downfall of the Ptolemaic world system. Copernicus tried to repre- 
sent his system with the help of the concepts of the Aristotelian 
scholastic philosophy. If, however, we read the dialogues of Galileo 
about the two world systems, we encounter a very different tone. 
Here the basic concepts of Aristotelian physics were taken up and 
examined. In the teachings of Aristotle and his school, concepts like 
“light” and “heavy,” “above” and “below,” “natural” and “forced” 
motion, which were usable only for a very limited domain of experi- 
ences, were made the basis of all theoretical physics. Galileo showed 
that it was just this use of concepts outside of their natural realm of 
validity that prevented the followers of Aristotle from understanding 
modern physics. I do not mean by this to belittle Aristotelian physics, 
which was an outstanding contribution for its time; I only wish to 
show that what was enlightening in Galileo’s writings was his setting a 
limit to the misuse of auxiliary concepts. And it is this protest against 
the misuse of merely auxiliary concepts in general philosophical proofs ^ 
that I consider to be an essential characteristic of enlightenment. 

Every period of physics has its auxiliary concepts, and every suc- 
ceeding period misuses them. Hence in every period a new enlighten- 
ment is required in order to abolish this misuse. When Sir Isaac Newton 
and his contemporaries made the concepts of absolute space and time 
the basis of mechanics, they were able to represent a large domain of 
physics properly and without contradictions. It does not follow, how- 
ever, that these concepts form a basis of mechanics that is satisfactory 
from the standpoint of the theory of knowledge. When Mach criticized 
the foundations of Newtonian mechanics and tried to eliminate absolute 
space from it, he was the direct continuer of the work of Galileo. For in 
absolute space we still had a remnant of Aristotelian physics. And 
when Einstein joined Mach and in his general theory of relativity really 
erected an edifice of mechanics in which space and time properly 
speaking no longer occurred, but only the coincidence of phenomena, 
the elimination demanded by Mach of the au^ary concepts of space 


73 



modem science and its philosophy 


and time— useful only in a limited domain— was completed. Einstem is 
the first tliinker to found a physics entirely free of Aristotelianism_, 

In the Age of Enlightenment proper I also see a struggle against the 
misuse of auxiliary concepts. If we leave out of the discussion the 
political and social aspects, then, theoretically considered, the criticism 
at that time was directed against the fact that theologic concepts, 
which were formed for dealing with certain psychic experiences of 
human beings, were made the foundation of all science throughout the 
Middle Ages and even at the beginning of the modern era. These con- 
cepts, no matter how appropriate they may be to restore hope and faith 
to the struggling human soul, are nevertheless only auxiliary concepts 
limited to this domain and are not suitable to be the epistemologic 
foundation of our knowledge of nature. This critical point of view 
emerged with great energy at that time. Today even theologians have 
adopted the view that the Bible is not a scientific textbook. Indeed, 
many Protestant theologians, proceeding still further in the direction 
of enlightenment, now teach that all theologic truths are only state- 
ments about inner experiences. 

The natural science of the Enlightenment also needed au,xiliary 
concepts for its development. In this way the concepts of matter and 
atom began to play a decisive role. All at once these auxiliary concepts , 
were being applied to everything in the world; materialism, as it is 
called, was bom. The fact that matter, too, was only an auxiliary con- 
cept was forgotten, and people began to regard it as the essence of the ' 
world. Soon criticism of this viCxv set in. But although this criticism of 
the misuse of auxiliary concepts usually serves only scientific progress, 
here it had an additional effect. Since the ideas of the Age of Enlighten- 
ment were not pleasing to the mling powers, the criticism of the misuse 
of enlightenment was used to discredit enlightenment itself. Because 
the rationalists misused auxiliary concepts, it was said of them that their 
protest against the theologic world picture was unjustified. This view is, 
of course, logically not tenable, for the fact is that their criticism did not 
go far enough. However, there are always thinkers who are so con- 
stituted that their thinking ultimately leads to the conclusions required 
by the mling powers. An attempt was made to overthrow enlighten- 


74 



Ernst Mach's philosophy of science 


ment by skepticism. Very appropriately Nietzsche says, of the part 
taken in this work by certain philosophers: 

The philosopher against his rival, e.g., science: now he becomes a 
skeptic; now he reserves for himself a form of cognition which he denies to 
the scientist; now he goes hand in hand with the priest so as not to arouse 
the suspicion of atheism, materialism; he regards an attack on himself as 
an attack on morality, virtue, religion, order— he knows how to discredit his 
opponent as a "seducer” and an “underminer”; now he goes hand in hand 
with the authorities.®^ 

In reality, however, only that part of enlightenment was refuted 
which was not enlightenment. Nevertheless, because of the weight of 
the authority of those involved, this disparagement of the great 
achievements of the eighteenth century has had considerable influence. 
There is perhaps no one among us who did not acquire a prejudice 
against enlightenment at school dining his youth. 

I admit, of course, that the great spirits of the Enlightenment, 
Voltaire, d’Alembert, and the rest, were imitated by many shallow 
writers who diluted their criticism more and more and descended to 
intolerable banality, ending up by continuing themselves the misuse 
of the auxiliary concepts. I admit, too, that this shallowness belongs 
to the essence of the Enlightenment. Once the misuse of the old con- 
cepts has been exposed, there is not much of an original nature left to 
say. The temptation to dull triviality is strong and the number of those 
who have fallen victim to it is great. 

Nevertheless, all this provides no evidence whatsoever against the 
philosophy of enlightenment. A man who has freed himself once and 
for all from the fears of the usual stigma of heresy will say that the 
task of our age is not to fight against the enlightenment of the eight- 
eenth century but rather to continue its work. Since that time there 
has occurred so much exaggerated application of entirely new auxiliary 
concepts which are useful in hmited domains that there is plenty of 
new work to be done. 

To this work Mach dedicated himself. He approved enthusiastically 
of the eighteenth-century enhghtenment. This does not mean, how- 

Friedrich Nietzsche, The WiK to Fotoer, No. 248. 

75 



modern science and its philosophy 


ever, that he began to idolize the eighteenth-century concepts like 
materialism. Rather, in him lived the spirit of those great men; it drove 
him on to protest against the misused concepts of his time just as they 
had fought those of their time. Among the concepts that he fought 
there happened to be many of the favorite concepts of eighteenth- 
century enlightenment. 

This is what I mean when I call Mach the representative of the 
philosophy of enlightenment of our period. Since his youth was lived 
during the time of materialism, it is no wonder that so many of his 
works are devoted to the struggle against mechanistic physics and 
atomistics. 

If we accept Mach’s attitude as that of an enlightenment phi- 
losopher it will be easier for us to understand many features of his 
teachings and many of their effects. In the first place, there is their 
strongly suggestive influence— one might even say their virulence— 
which, in spite of many contemptuous judgments by professional phi- 
losophers, commands attention. Study calls the positivism of Mach 

a still completely unsatisfied existence, a kind of philosophical beast of 
prey, hungry for a victim.” 

As in the case of the philosophers of the enlightenment, so also in the 
case of Mach, it developed that the disciples and followers displayed 
an excessive tendency towards shallowness. Furthermore, to Planck’s 
criterion of the fruits the present opinion provides an answer: the fruits 
of Mach’s teachings are not the writings of his physical and philo- 
sophical followers, but rather the enlightenment of minds brought 
about by them— a fact which even Planck recognizes. By this I do not 
mean to assert that Mach has no importance in other respects. I do 
think, however, that this is the best summary of his position in the 
general intellectual life of our times. 

In this opinion I am strengthened also by the striking agreement 
of his views with those of a thinker for whom he cannot have had any 
great sympathy, Friedrich Nietzsche. This agreement was first pointed 

” Study, op. ctt., p. 24. 


76 



Ernst Mach's philosophy of science 

out by Kleinpeter.'^ The more one delves into the posthumous works 
of Nietzsche, the more clearly one observes the agreement, particularly 
in the basic ideas related to the theory of knowledge. Now Nietzsche 
is the other great enlightenment philosopher of the end of the nine- 
teenth century. The harmony of his epistemologic views with those of 
Mach, who had gone through an entirely different course of instruction 
and whose temperament and ethical ideals were entirely different, 
seems to me to be evidence for the fact that such views must have 
penetrated to the enlightened minds of that time. 

That great master of language, Nietzsche, formulated these ideas 
with extraordinary force and impressiveness when he said: 

I behold with amazement that science today has resigned itself to being 
relegated to an apparent world; a true world, be it what it may: in any case, 
we have no organ of cognition for it. Here one might ask: by what organ of 
cognition is one led to assume this antithesis? . . . From the fact that a 
world which is accessible to our organs is also understood to he dependent 
on these organs, from the fact that we understand a world as being sub- 
jectively conditioned, it does not follow that an objective world is in any 
case possible. Who prevents us from thinking that subjectivity is real, 
essential? The “in itself” is a contradictory conception, to be sure: a "quality 
in itself” is nonsense: we have the concept “being,” “thing,” always only as 
a relation concept . . . The bad part of it is that’ along with the old 
antonyms “apparent” and “real" the correlative judgments of value have 
been propagated: “of little worth” and “of absolute worth" . . 

Elsewhere Nietzsche says: 

That things have a quality in themselves, quite apart from any in- 
terpretation and subjectivity, is an idle hypothesis: it would presuppose 
that to interpret and to be a subject are not essential, that a thing detached 
from all relations is still a thing.^’ 

Nietzsche’s most significant expression of the positivistic world 
conception is probably given in the aphorism, called “On the Psychol- 
ogy of Metaphysics,” where he attacks with cutting sharpness the 
employment of very frequently misused concepts: 

“H. Kleinpeter, Der Fhenomenalismus (Leipzig, 1913). 

“ Nietzsche, The WUl to Potver, No. 289. 

“ Op. cit.. No. 291. 


77 



modern science and its philosophy 


This world is apparent: consequently there exists a true world;— this 
world is conditional: consequently there exists an unconditional world;— 
this world is full of contradictions: consequently there exists a world that is 
free from contradictions;— this world is changing: consequently there exists a 
permanent world;— all false conclusions: (blind faith in the reasoning: if 
there is A, there must also be its antithetical concept B ) 

It is not to be denied that the philosophy of enlightenment possesses 
a tragic feature. It destroys the old systems of concepts, but while it 
is constructing a new system, it is also already laying the foundations 
for new misuse. For there is no theor)' without auxiliary concepts, and 
every such concept is necessarily misused in the course of time. The 
progress of science takes place in eternal circles. The creative forces 
must of necessity create perishable buds. They are destroyed in the 
human consciousness by forces which are themselves marked for 
destruction. And yet, it is this restless spirit of enlightenment that keeps 
science from petrifying into a new scholasticism. If physics is to be- 
come a church, Mach cries out, I would rather not be called a physicist. 
And with a parado.xical turn, Nietzsche comes out in defense of the 
cause of enlightenment against the self-satisfied possessor of an en- 
during truth: 

The assertion that the truth is here, and that an end has been made of 
ignorance and error, is one of the greatest seductions that there are. Assum- 
ing that one befieves it, then the wiU to test, investigate, predict, experiment, 
is crippled: the latter can itself become wanton, can doubt the truth. The 
“truth” is consequently more ominous than error and ignorance because it 
binds the forces with which one can work for enlighteiunent and knowl- 
edge.” 

Of these forces, however, at the turn of the century Mach was one 
of the mightiest. 

“ Op. dt.. No. 287. 

” Op. dt.. No. 252. 


78 



CHAPTER 



Ernst Mach and the unity of science 

I N the year 1882 the famous American philosopher and psychol- 
ogist William James made a tour through Europe and every- 
where met scientists interested in his field of work. At the end of 
October, James came to Prague and met Ernst Mach. James described 
his impressions of this meeting in a letter he wrote to his wife in 
America: 

Mach came to my hotel and I spent four hours walking and supping 
with him at his club, an unforgettable conversation. I don’t think any one 
ever gave me so strong an impression of pure intellectual genius. He ap- 
parently has read everything and thought about everything, and has an 
absolute simplicity of manner and a winningness of smile, when his face 
lights up, that are charming. 

I do not want to speak here about the wide activity of Mach in 
physics, physiology, psychology, and the history and logic of science, to 
do which would necessitate a series of papers. Instead, I shall speak 
about Mach’s activity only in so far as he may be considered one of 
the spiritual ancestors of the Unity of Science Movement and, par- 
ticularly, the real master of the Vienna Circle. 

A hundred years after the birth of Ernst Mach we must not forget 
or underrate the fact that he is still very much alive today. Surveying 

the opinions of the scientific workers of today, we find many of them 

* • 

79 



modern science and its philosophy 


who decidedly reject his doctrine. On the other hand, there are a 
great many scientists who enthusiastically express their full agreement 
with him. But among the scientists who are acquainted with Mach’s 
doctrine there are very few whose behavior toward him is neutral or 
indifferent. In spite of this fact, opinions about what are the most 
characteristic features of his doctrine are very different and sometimes 
even contradictory. 

On the one hand, Mach is described as the most radical opponent 
of every attempt to introduce into science factors that have any tinge 
of spiritual tendency. He saw even in a concept which for ages has 
been in physics as usual as the concept of force a detrimental remainder 
of the obsolete world picture of primitive man, which was animistic 
and fetishistic. On the other hand, we are told that according to Mach 
our world consists entirely of perceptions or complexes of perceptions; 
that there is no such thing as matter for the building of a world. For 
this reason Mach has been proclaimed as a champion of the idealistic 
philosophy within modem science and a leader in the struggle against 
materialism. 

To hint at another difference of opinion: on the one hand, Mach is 
said to claim that the sole function of science is to record observable 
facts and to sum them up by economical and suitable formulas. The 
scientist has, according to Mach, to be on his guard against daring 
generalizations, through which an animistic or metaphysical element 
may insinuate itself into science. On the other hand, physicists working 
in their research laboratories accuse Mach of not recognizing the 
existence of objective facts. According to these people, Mach main- 
tains that there are only subjective opinions of single physicists, but no 
real facts; there is no real physical world, the exploring of which was 
supposed to be the goal of the research work of the physicists. For this 
reason Mach’s doctrine is even alleged to have a paralyzing influence 
on research work and hence on the progress of science. For only the 
belief in a real, an objective, world can give to the physicist the mental 
activity and force needed for his diJBBcult achievements. 

"Whence these different judgments, sometimes even contradictory 
to each other, about the chief lines of Mach’s philosophy? Why is the 


80 



^ Ernst Mach and the unity of science 

essence of Mach’s doctrine described by different authors in such 
different ways? The chief reason for these differences is, I think, that 
philosophers, and sometimes scientists too, endeavor to discuss Mach’s 
doctrine in the language of traditional philosophy. In this language 
such terms occur as “idealism,” “spiritualism,” “materialism,” “real 
objective world,” “subjective opinion of the real world.” But the point 
is that it is impossible to describe Mach’s doctrine in this language, im- 
possible to describe it at all in terms of traditional philosophy. If we 
want to form an adequate conception of Mach’s doctrine, we must 
never forget that he always declined the title of philosopher. Mach has 
been praised by a great many philosophers for this modesty. But I do 
not think that it was exactly modesty. He wanted rather to draw a 
conspicuous boundary line between his own doctrine and the doctrine 
of traditional philosophy. 

Of first importance, it seems to me, for an understanding of the chief 
line of Mach’s thought is a passage in the introduction of his book, 
“The Analysis of Sensations,” ^ in which he explains his chief aim in 
writing the papers that are commonly described as philosophic. Mach 
starts from the fact that scientists are accustomed, each in his special 
field, to make use of a certain system of concepts, or, to put it more 
exactly, of certain technical terms, a certain technical language which 
is very suitable within this special field, say physics. But this special 
technical language may become very unsuitable and even misleading if 
it is applied to the description and formulation of the frontier problems 
which arise when we pass from one special science to a neighboring 
science, say from physics to biology or to psychology. Mach’s own 
words are: 

1 do not claim the title of a philosopher. I want only to take, in physics, 
a standpoint which does not have to be abandoned immediately when we 
look over into the field of another science. For all the sciences ultimately 
form a whole. What I am stating, I am perhaps not the first to state. 
Furthermore, I do not want to bring forward my explanation as an ex- 
traordinary achievement. I think, rather, that the same line would be taken 
by anyone who tried to survey a field of science that is not too narrow. 

^E. Mach, Beitrage xur Analyse der Empfindungen (Jena; G. Fischer, 1886; 
English translation, Chicago, 1914). • 


81 



modern science and its philosophy 


According to Mach this desire to make use of a unified mode of ex- 
pression in all fields of science is a consequence of the economical 
design of science. This design implies the comprehension of as many 
facts as possible by the simplest possible system of propositions. 

Since Mach dealt in his papers with so many different kinds of 
problem in the fields of physics, physiology, and psychology, a great 
many scientists did not discover what the chief trend of these papers 
was. In order to find the clue to it, we have to read attentively another 
passage in the same introduction. Mach says there expressly: 

The basis of aU my investigations into the logical foundation of physics 
as well as into the physiology of perceptions has been one and the same 
opinion: that all metaphysical propositions must be eliminated, because they 
are idle and disturbing to the economical design of science. 

And Mach’s famous book “Mechanics and its Development’’” begins 
with the sentence: 

"The tendency of this paper is an elucidatory or, to put it more ac- 
curately, an anlimetaphysical one.” 

In the reports on Mach’s teaching by advocates of traditional phi- 
losophy, one may often read that Mach’s chief doctrine was that the 
world consisted of perceptions and not of material particles. But from 
the passages that I have just quoted, in which Mach expressly describes 
the chief aim of his investigations, it is clear that these investigations 
were quite unconcerned with such problems as whether the world 
consists of perceptions or of matter. This is rather the typical way in 
which traditional philosophy likes to put a problem. It is just this way 
of putting a problem that Mach emphatically rejected. 

Hence Mach’s chief tendency may be described by the phrases, 
“unification (that is, economical presentation) of science” and “elimina- 
tion of metaphysics.” We shall find that these two aims are very closely 
connected with each other. And we shall see that the most widely 
popularized doctrine of Mach, according to which the real world 
consists of perceptions, was never formulated by him in this meta- 

®E. Mach, Die Mechanik In ihrer EntuHckelung (Leipzig: F. Brockhaus, 1883; 
English translation, Chicago, 1893). 


82 



I 


Ernst Mach and the unity of science 


physical way. What was really in Mach’s mind when he maintained 
that our world is built up of perceptions or complexes of perceptions 
was not that this phrase “built up of perceptions” is in any sense a 
statement concerning a property of the real world, but only that it is j 
a useful means for the unification of science and the elimination of 
metaphysics. It would be a misunderstanding of Mach’s aims to believe 
that the construction of the world by perceptions, which was merely 
a means to an end, was the true end of his philosophy. Many of his 
philosophical interpreters stick fast at this means to an end, “the per- 
ception language,” and neglect the real purposes of Mach’s doctrine, 
the unification of science and the elimination of metaphysics. For it is 
just by the presentation of some scientific results in terms of meta- 
physics that the unification of science is, according to Mach, gravely 
imperiled and sometimes even frustrated. 

If one describes physics as the science of matter, biology as the 
science of life, psychology as the science of the mind, sociology as 
the science of the collective mind, metaphysical concepts or words such 
as “matter,” “life,” “soul,” "collective soul,” are introduced, and it is 
obvious that words like “matter” and “soul,” for example, are probably 
not reducible to the same terms. It is easy to prove that the introduc- 
tion of expressions of this kind renders impossible the representation of 
our experiences by a unitary system of terms; in other words, it renders 
the unification of science impossible. 

^To remove these diflSculties Mach suggested the formulation of 
the laws of physics as functional connections among perceptions such 
as “green,” “warm,” “hard,” including also, of course, the space and 
time perceptions. Every physical experiment consists in observing how 
the alteration of some perceptions is connected with the alteration of 
others. If no perceptions touching om: own bodies intervene— for ex- 
ample, if there is no altering of perceptions by intoxication of our 
nerves— we are in the field of physics. If we observe the connections 
between perceptions, including perceptions arising from alteration in 
om: own bodies, we are in the fields of physiology or of psychologyf^ 
But it is clear that we can no longer prove the impossibility of a unified 
language of science if we start from Mach’s perception language in- 

83 



modern science and its philosophy 


stead of the metaphysical terminology of traditional philosophy, and, 
we must admit, sometimes even of traditional science. 

On the contrai)/,% we start from Mach’s standpoint and formulate 
all scientific propositions in terms of perceptions, the unification of 
science becomes possible. Mach never maintained that our world 
consisted of complexes of perceptions, but that every scientific proposi- 
tion was a statement about complexes of perceptions. Whether it be 
a proposition of physics, biology, or psychology, it can only be proved 
or refuted by comparison with observation.'^ 

•According to Mach, the unification of science is possible, but only 
by formulating all scientific propositions as propositiras about com- 
plexes of perceptions, in the widest sense of this word. 'Every proposi- 
tion that states something about our observations contains as predicate 
some term like “green,” “warm,” “joyful,” “painful,”— perception terms, 
as Carnap calls them. A proposition that is not reducible to propositions < 
containing only perception terms as predicates cannot be checked by I 
experience; it is a metaphysical proposition. Hence to Mach the ex- 
pression “elimination of metaphysics” means the elimination of all sen- 
tences that are not reducible to sentences containing only perception 
terms as predicates’? The elimination from science of metaphysical 
propositions leaves only sentences of a homogeneous type, namely, 
sentences with perception terms as predicates. Therefore, if we demand 
of science an economical representation of our experiences, that is, a 
representation by a unified system of concepts, we must admit only 
propositions that are reducible to propositions containing only per- 
ception terms as predicates. 

\This is the real meaning of Mach’s doctrine that all propositions 
of science deal with perceptions. He did not want to make a statement 
about the question of what the world consists of,| but he wanted to 
point out how the propositions of science had to be formed in order to 
make possible a unification of science. His result is this: The unifica- 
tion of science is not possible except by the elimination of meta- 
physical propositions.Vrhen only propositions of a homogeneous type 
remain. Hence we can form from them a coherent logical system.\ 

The elimination of metaphysics from science was for Mach, as we 


84 



Ernst Mach and the unity of science 


now understand, not a demand arising from some antimetaphysical 
mood, but the only means of making possible the unification of science. 
According to Mach, metaphysics must be eliminated because it is con- 
tradictory to the economical function of science. 

A great many people were puzzled because Mach’s philosophy, 
which was supposed to be a sort of idealism (similar to the philosophy 
of Bishop Berkeley), changed so easily, or— to speak in terms of ideal- 
istic philosophy— degenerated so easily, into physicalism. We have seen 
that even the Vienna Circle went over very quickly from the phe- 
nomenal language used, following Mach, by Carnap as well as by 
Schlick, to the physical language claimed by Neurath. In physicalism, 
which now plays a great role in aU papers that start from the viewpoint 
of logical empiricism (in the most coherent and exact way in the 
papers of Carnap), a language is used that seems very near to ma- 
terialism. 

So it has been for a great many philosophers a riddle and almost a 
source of irritation that the opinions within a group claiming to possess 
a particular sense of logical consistency could oscillate so easily be- 
tween the extreme poles of human thinking, which idealism and 
materialism are supposed to be. 

But this antithesis, which, according to traditional philosophy, exists 
between materialism and idealism is not, according to Mach, a scientific 
antithesis. Mach disliked the use of terms such as “idealism” and 
“materialism” and if he did use them it was only to reject them. 
Though he rejected materialism as well as idealism, this rejection does 
not mean that he tried to take a mediating standpoint between them. 
/Foi him, both idealism and materialism are systems of metaphysical 
propositions, not scientific theories. For they can be neither proved nor 
refuted by expeiienc^Mach had what one might call the instinctive 
aversion of a genuine scientist to the use of vague terms like “idealism” 
or “materialism” in science. This aversion sometimes induced bim to 
make statements against one or the oBier metaphysical system. And 
his statements were often misinterpreted as statements on behalf of 
the other system, just as if, by rejecting one sort of metaphysics, he 
advocated the contrary metaphysics. From Mach’s point of view the 


85 



modern science and its philosophy 


question of “idealism” or "materialism” cannot be put as a real scientific 
problem. Every attempt to exploit the achievements of science to 
bolster up idealistic or materialistic metaphysics is from the beginning 
doomed to failure. 

What Mach felt instinctively we can today formulate in words, if 
we take the standpoint of logical empiricism, as it was formulated 
very precisely in Carnap’s book “Logical Syntax of Language”* and 
in his paper “Testability and Meaning.” * 

The transition from the supposed idealistic to the physicalistic 
conception of science took place so smoothly within the Vienna Circle 
because^according to the doctrine of logical empiricism, the question 
to be put was not whether idealism and materialism were right opinions 
about the real world, but only what language, phenomenal or physical, ^ 
was the more suitable for giving an economical and unitary account of 
our experiences. Since either of these languages may be more suitable 
within a limited field than the other, the choice of an accepted language 
has nothing to do with the question whether our real world consists of 
perceptions or of matter. What is of essential importance is only the 
question whether we believe that it is possible to comprehend all fields 
of science in one and the same language. If the unification of science, 
in this sense, is held to be possible, as Carnap and the adherents 
Unity of Science Movement maintain, it is of secondary importance 
whether this unification be achieved in terms of perceptions, as Mach 
believed and Carnap proved to be right in his first paper, “The Logical 
Construction of the World,”* or whether the physical language is to 
be introduced, as Carnap proclaims in his recent papers in accordance 
with Neurath’s suggestions.'^ 

The essential alternative for our conception of science is rather this: 
Do we, in accordance with traditional philosophy, maintain that the 
question whether the real world consists of matter or of perceptions, 
and other questions like it, are scientific questions, or do we, vrith 


®R. Camap, The Logical Syntax of Language (New York: Harcourt, Bract 
1937). 

* Philosophy of Science (1936 and 1937). 

®R. Camap, Der logiscke Aufbau der Welt (Berlin, 1928). 


86 



Ernst Mach and the unity of science 


Mach, eliminate metaphysical questions of this sort from science as 
disturbing its economical character and put the question just mentioned 
in the way in which it is put by logical empiricism? 

Then we ask, what language is the most suitable as the language of 
a unified science. From this point of view, the metaphysical question 
seems to be, as Mach expressed it, an idle one. And the question as logi- 
cal empiricism puts it, whether the phenomenal language or the physi- 
cal language is more suitable as the language of unified science, ceases 
to be a question of profound metaphysical importance and becomes a 
question of convenience. It is perhaps comparable to the question of 
what system of symbols is the most suitable for the introduction of a 
unified symbolism into logic. 

If we want to describe Mach’s role in the history of human thought 
and in the development of science in a particularly comprehensive 
and impressive way, we can do it, it seems to me, by a clear-cut 
antithesis. The traditional conception of science is connected with a 
certain opinion about the importance of metaphysics to science. Ac- 
cording to this conception, there are two methods of science: 

(1) The first method is restricted to recording facts, including 
empirical rules summarizing facts. The adherents of this method of 
scientific activity take care not to introduce sweeping generalizations 
and hypotheses, because by them a metaphysical element may easily 
creep into science. This kind of scientific activity possesses, according 
to its adherents, the advantage that all propositions admitted by science 
are secured by experience or logic and are clear and intuitive. The 
scientists of this group are anxious not to introduce vague phrases. But 
by proceeding in this cautious way only special sciences of small extent 
are created. Physics and biology and psychology have each collected a 
lot of facts and rules, but there is no link between these special de- 
partments. This conception of science is often referred to as the 
“positivist” conception, it is not at all in accordance with the conception 
of the so-called “logical positivism,” neither does it agree with Mach’s 
r doctrine. This pseudo-positivist conception of science does not satisfy 
our aspirations toward the unity of science. 

(2) Therefore, beside this positivist conception, or more exactly, 

• 

87 



modern science and Hs philosophy 


over it, the metaphysical conception of scientific activity has always 
existed, because it has been supposed to be better adapted to satisfy- 
ing our desire for a synthesis of knowledge and the unity of science. 
According to this conception we can reach the unity we are striving 
after by the introduction of daring metaphysical generalizations and 
hypotheses. By means of these metaphysical generalizations the 
separate sciences can be summarized and unified into a unitary science. 
The general principles of this unified science are, of course, metaphysi- 
cal propositions. The most famous metaphysical system, which was sup- 
posed to comprehend and represent all special sciences, was the system 
of Hegel. In this system all separate sciences, such as mathematics, 
physics, biology, are presented as steps in the self-evolution of the 
absolute spirit. An example of the sort of metaphysical proposition 
that serves to achieve the unification of science is Hegel’s fundamental 
theorem of dialectics, namely, “every quantity, if sufficiently increased, 
turns into a quality.” This theorem is supposed to be valid in physics 
ns well as in biology, in biology as well as in history. It is particularly 
interesting because it continues to play a great role and is not confined 
to the adherents of Hegel’s idealistic metaphysics. This theorem and 
many like it are taken over into dialectical materialism, the official 
philosophy of the Soviet Union of today. By means of propositions of 
this kind the barriers between the separate sciences are broken down 
and the unification of science is achieved, but only at the price of in- 
troducing large systems of very vague propositions. With regard to 
such metaphysical propositions a general agreement among scientists 
would never be reached. 

’The opinion of traditional philosophy and traditional science has 
been for ages that scientific activity has only these alternatives: either 
that only theorems that can be proved by experience or logic are 
recognized as legitimate in science, in which case the separate sciences 
remain separated from each other by insurmountable barriers, or that 
we admit the introduction of metaphysical propositions, in which case 
the unification of science can be achieved, but we have to deal with 
propositions that will never be recognized by all scientists. 

To put it still more concisely; either the renunciation of the unifies- 


88 



Ernst Mach and the unity of science 

tion of science or the introduction of metaphysical propositions into 
science. 

The great importance of Mach's activity lies in the fact that he 
declined to recognize these alternatives. He proclaimed, rather, the 
unification of science by means of the elimination of metaphysics. 

It is just this phrase that is the clue to the imderstanding of Mach’s 
doctrine, and of his papers, which seem to deal with so many sub- 
jects and such difiFerent fields of science. What always mattered for 
Mach was the opportunity to accomplish the program that we have 
just outlined. And it is just this program of Mach that we may adopt 
as the program of our “Unity of Science Movement,” of our Congresses 
and of our Encyclopedia. If Mach’s centenary has been celebrated by 
so many physicists, physiologists, psychologists, and historians of 
science we may take a special pride in these celebrations, in which we 
have a special right to honor him as one of the spiritual ancestors of 
the “Unity of Science Movement” For, I think, within our movement, 
the harvest of the seed scattered by Mach is particularly rich and in die 
strictest accordance with his true intention. 


89 



CHAPTER 



physical theories of the twentieth century 
and school philosophy 

% 1 #HAT is the significance of present-day physical theories 

\i\m for the general theory of knowledge? There are many physi- 
Y ycists and philosophers whose answer would be, “Nothing.” 
On what grounds many philosophers give this answer, I cannot and 
do not wish to investigate here. How does it happen, however— and 
this question will ser\e as our starting point— that so many physicists 
assert that the greatestuavol ntions in t he theorie s of physics are 
.^^papgKlp nf pVia nging th fi principles ot the general theory of knowl- 
edge? For example, one finds in physical works on the theory of 
relativity the thesis, often defended with passion, that the relativistic 
revision of the measurement of space and time has no “philosophical” 
consequences. 

Anyone who has occupied himself to some extent with the historical 
development of physics will be struck at once by the similarity of this 
thesis to what took place during the period of those great revolutions 
in the theories of physics that led from the medieval, scholastic con- 
ception of nature to the conception prevalent today, revolutions as- 
sociated above all with the names of Copernicus, Galiledj KeplefTi'hus, 
one reads that the adherents of the heliocentric theory— at that time 
considered revolutionary— contended most zealously that the Copemi- 

90 



^twentieth-century physics and school philosophy 

can “revolution” had brought about something that was o nly mathe - 
matidallv ai^d physic ally new, that it had made no change at all in the 
general philosophic conception of the world. At the famous trial of 
Galileo, when pressure was brought to bear on him to recant his 
doctrine, it was not a question of his swearing that he no longer be- 
lieved in the motion of the earth, as one often reads in superficial 
presentations, which have immortalized the famous misquotation: 
“Nevertheless, it does movel” What the Inquisition actually wanted of 
Galileo was only that he confess that the doctrine of the motion of 
the earth was correct merely as a mathe matical &tion, but was false 
as a philosophical doctrine. We can also find in the standpoint of the 
Inquisition something corresponding to the modern relativistic co ncep- 
tion. According to the latter, we cannot say that “in reality” the earth 
moves and the sun stands still, but only that the description of phe- 
nomena turns out to be simpler in a coordinate system in which this is 
the case. More than this was demanded of Galile o, however. He was 
expected to admit that the heliocen tric co nce p tion was a mathematical 
fiction, whereas the geocentric one ytas .a. pMosaphieal ..truth. It is 
readily seen that even this standpoint of the medieval authorities finds 
its analogue in our times. Today, too, a fictionalistic conception is 
often presented for the purpose of bringing out more strongly, by con- 
trast, the “eternally vali d.” “philosophically reasonable” truths. For 
instance, it is very often asserted by philosophers, and sometimes also 
by physicists, that non-Euclidean geometry and the Einstein measure- 
ment of time are piathematicaLfictions, whereas Euclidean geometry 
and absol ute time are, in the very nature of things, established truths. 

Still more often we find, however, that ^physicists refuse to make 
any decisions concerning such questions as time, space, causality, pre- 
ferring to leave these for the competent specialist, the philosopher or 
the e;psl£0}plo.gist. Since today no one need have the same fears as 
Galileo, this refusal must arise from a conviction that can be formulated 
roughly as follows: “There are questions that are so profound that 
they cannot be solved by the exact sciences.” In connection with this 
point, some believe that there is a special method, the “philosophical,” 
with the help of which such questions as those concerning the nature of 


91 



modern science and its philosophy 


time, space, and causality can be answered, while others regard these 
problems as forever insoluble, as “eternal riddles.” 

The classical expression of this resignation of the exact sciences 
was the famous speech of Emil du Bois-Reymond which dates back 
to 1872 and bears the title “On the Limitations of Natural Science.” ^ 

This speech, which culminates in the declaration “Ignorabimus”—v/e 
shall never know— has been quoted countless times— triumphantly by 
the belitders of the scientific world conception, with melancholy assent 
by its adherents. The speech in its essentials has been accepted by most 
philosophers and scientists as irrefutable truth. In the history of the 
scientific world conception it has been the scientist’s tri p to Canos sa. 
If we reflect by what arguments Du Bois-Reymond arrived at his 
ignorabimus, we must reach the conviction, considering the present 
state of the epistemology of the exact sciences, that it is high time to 
unroll the question once more, and once again to see whether the 
d§spamng point of view concerni ng scientifi c , cognition is really 
unavoidable. 

Du Bois starts with the thesis: 

The cognition of nature is the reduction of changes in the material 
world to motions of atoms, acted upon by central forces, independent of 
time ... It is a psychological fact of experience that wherever such a 
reduction is successfully carried through our need for causality feels satisfied 
■for the time being. 

There still remains, however, the question of how matter is able to 
exert the central forces. This question naturally cannot be reduced 
again to central forces. To quote Du Bois further: 

No one who has given any thought to the subject can fail to recognize 
the transcendent nature of the obstacles that are here before us . . . Never 
shall we know any better than today what it is that haunts the space where 
matter is found. For even the spirit of Laplace could not be any wiser on 
this point than we . . . Our cognition of nature is therefore enclosed by 
the two boundaries set up— on the one hand, by our inability to comprehend 
matter and force, and on the other hand, by our inability to understand 
mental processes in terms of material conditions. 

^ Uber die Grenzen de^ Naturerkennens, 

92 



% twentieth-century physics and school philosophy 

If we disregard the problem of the connection between the spiritual 
and the material, since it does not concern us here, Du Bois then sees 
the limits of our knowledge of the world primarily in the i mpossibility 
of understanding the nature of matter and force. He continues: 

Within these limits the scientist is lord and master; he dismembers and 
builds up; . . . beyond these limits he cannot and will never be able to go. 
With respect to th e riddles of th e material world . . . ignoramus; . . . 
with respect, however, to the riddle of what are matter and force, and how 
they are able to think, he must decide, once and for all, on a verdict much 
more di£Scult to rend er: iqnorabimus . 

But what does it mean when we say that a question is insoluble? 
Let us suppose, for example, that someone has asserted that the 
problem of a regular transport route to the planet Neptune is insoluble, 
or that the production of a living organism from lifeless matter is im- 
possible. Despite this assertion, the person making it can describe quite 
accurately the concrete experience we should have if the problem were 
solved. We cannot, however, in any way— not even approximately— 
imagine what we should have to experience in order to be able to say 
that the problem of the nature of matter or force has been solved, or 
that one knows, as Du Bois requires, “what it is that haunts the space 
where matter is fornid.” When, for example, Heinrich Hertz, as is often 
said, elucidated the nature of light, this was not at all in the sense 
meant by Du Bois. According to Hertz, light and electromagnetic phe- 
nomena are associated with the same equations but have different 
wavelengths; the nature of light is thereby made no clearer than it was 
before, since the nature of electricity is also in this sense an eternally 
insoluble riddle. 

Thus there are two types of unsolved and perhaps insoluble prob- 
lems, which Du Bois characterizes by means of the words ignoramus 
and ignorabimus. Anyone who is used to working on the real solutions 
of problems will feel a certain displeasure when he turns to problems of 
the second kind. For he is accustomed to proceed with the solution by 
first picturing to himself the experience corresponding to the completed 
solution, and then working until he succeeds in bringing about the 
realization of the desired experience. If, however, we cannot state what 


93 



modern science and Its philosophy 


this experience is to consist in, have we really set up a problem? 

As a matter of fact, we find very often that the physicist, as physicist, 
declines to work on problems set up in this way; yet in another comer 
of his soul he admits that such problems might be attacked with other 
methods— not physical, but “philosophical,” as they are called. In seek- 
ing the reason why physicists— who as such attach the greatest value to 
exact formulations of questions— in spite of their displeasure admit the 
possibility of something quite different, I believe it is necessary to take 
into account the fact that many physicists, when not working in their 
own fields, are inclined to a world conception that has become rooted 
in the educational system through centuries-old tradition, a conception 
which w'e will call simply the world conception of “school philosophy.” 

We do not wish to investigate here the question of why so many 
physicists adhere to this school philosophy, in spite of the fact that it 
was just the critically thinking physicists who contributed most to its 
shaking; for the causes of this adherence are to be understood only 
psychologic ally, and perhaps s ociologicall y. Rather, we wish to find out 
what the standpoint of the school philosophy consists in, and how it 
has made so many scientists assent without opposition to the resigned 
^gnorabitnus. 

It is said that the philosophic schools are so far apart from one an- 
other in their points of view that it is not possible to perceive among 
them anything like a unified conception_oiJhe wojld. In spite of these 
individual differences of opinion, however, I believe that we can see 
today a common nucleus, which has been handed down through the 
centuries and, to a certain degree, has crystallized. Along with this 
there has been developing, at first timidly, then more and more boldly, 
but even now still quite cautiously, a new world conception, one that 
is gradually gaining strength with the progress of the exact sciences. In 
order to introduce some name, just as we called the traditional doc- 
trine the “school philosophy,” we will call the new one the “scientific 
world conception,” to indicate briefly that it recognizes no other 
cognition than the scientific. 

The school philosophy, whether it calls itselLxeal ism or i dealism, is 
characterized by the possession of a certain conception of what is 


94 



Jwentieth-century physics and school philosophy 


n alifid truth , hence also of what can be regarded as the real formula- 
tion of a problem. The basic ideas of this doctrine of the school phi- 
losophy cannot be presented better than by Henri Bergson’s introduc- 
tion to the French translation of the American psychologist, William 
James's, Pragmatism. Bergson says: 


For the ancient philosophers there existed a world, raised above space 
and time, in which all possible tmihsJiad dwelt since eternity. According 
to these philosophers, the truth of human judgments was measured by the 
degree to which they were faithful copies of those eternal truths. The 
modern philosophers, to be sure, have brought down tru th from heaven .t o 

would be a law governing the facts: if not ruling over them, at least ruling 
in their midst; a law actually contained in our experience; it remains for us 
only to extract it from the latter. Even a philosophy like that pf Kan t, which 
assumes that every scientific truth is such only in relation to the human 
mind, considers the true propositions as given a p riori bv human experie nce. 
Once this experience, in general, is organized by human thought, the 
whole work of science consists in breaking through the obstructing husk 
of facts, in the interior of which the truth is housed like a nut in its shell. 


earth, but they still regard it as son ^ — , , ^ 

ments. A proposition such as “Heat expands a body,” according to them. 


✓ 


It is readily seen that this conception of truth permits every type 
of question. It makes it difficult, however, to distinguish between 
sensible and meaningless formulations of problems. For to every ques- 
tion the answer can be found behind the h us k of fa cts if one bores 
with sufficient energy. In principle, it might then be possible to answer 
even such questions as those concerning the nature of matter and 
force. If, however, the shell of the nut is so hard that it can never be 
bored through, so that the answer cannot be extracted, we call the 
question “eternally insoluble,” and say resignedly . Ignorabimus . If 
we have this conception, we can also pose such questions as those that 
are most characteristic of the school philosophy: whether the outer 
world r eally exist s, and whpthf»r wr ran Imnw ths world in its t rue 
properties. To these questions the realist r eplies in the affirmative, the 
i dealist, in the negative. Neither can adduce any concrete experience 
as decisive for his answer. Both agree, however, that su ch a giiesHnn is 
a sensible problem. 


95 



modern science and its philosophy 


, There is no doubt that this standpoint of the school philosophy 
makes very diflBcult the acceptance and understanding of present- 
day p hysical theo rips. For example, from this point of view one can 
raise the question of what the “real” length of any body is. If the 
[theory of relativi ty ascribe s to a body different lengths with respect to 
•different reference systems, the adherent of the school philosophy will 
opine that this difference arises from “perturbations” of the measuring 
instrument, which make the “correct” measurement impossible in 
practice. This, however, does not prevent one length from being the 
“real” one as distinguished from the merely “apparent” measured 
lengths. Of the f amily o f reference systems moving uniformly and 
rectilinearly with relation to one another, only one can be really at 
rest, according to this conception. Since according to the relativity 
theory— and so far it has not been disproved experimentally— it is not 
possible by any experiment to determine which is th e system that is 
really at re st, for the adherent of the school philosophy this “being 
really at rest” is a fact that cannot reveal itself in any concrete ex- 
perience that man can have. 

One who considers it obvious that an electron must have at every 
instant a d efinite positiPiL-and—velocity— though the measurement of 
them may be impossible— finds it difficult to understand the basic 
principles of .quantum mecha nics. He is forced to interpret the 
quantum-mechanical calculations, which he uses nevertheless, in 
such a way that these definite positions and velocities of the electron 
do not determine its future. Since, on the other hand, the doctrines 
of the school philosophy in the field of mechanical phenomena re- 
quire strict detennioism, he is forced to assume for the motion of 
the electron some mystical vital causes, similar to organic life. True, 
this conclusion will be pleasant and congenial to some people; but I 
do not believe that it is useful for physical investigation. It does not 
follow— as many believe— from the theories of modern physics; it arises 
only from the desire to bring these new theories into harmony with the 
world conception of the school philosophy. 

The objection will perhaps be raised that in most of their investiga- 
tions physicists are not at all concerned about philosophy, and hence 


96 



^twentieth-century physics and school philosophy 


that the school philosophy cannot be a hindrance to the understanding 
of relativity or the quantum theory. Physicists examine these theories 
from a “purely physical” point of view and do not know anything at 
all about the philosophical conception of the world. If one studies 
closely the reactions of physicists toward the modem theories, how- 
ever, it will be found that the less they are accustomed to thinking 
about philosophical questions, the more their thinking is dominated 
by the traditions of the school philosophy. Experience has also shown 
that those physicists who declared, for instance, that the relativity 
theory was nonsense often spoke in the name of “pure, empirical sci- 
ence, free from speculation,” but took their arguments chiefly from the 
school philosophy, not from empiric ism. It need not be supposed that 
one has to make any philosophical studies in order to become ac- 
quainted with this world conception. In all knowledge that has come 
to us from the elementary school, in all metaphors of our language, 
it is implicitly contained. Its presence is not noticed because traditions, 
centuries old, make us take it for granted. The pure “ empiric ist” uses 
it under the name “common sen se.” Hence it is no wonder that it is 
just the physicist opposed to speculation who is easily inclined to the 
ignorabimus of Du Bois-Reymond, with his surrender of the scien- 
tific conception of nature. 

There are some physicists who do not understand the “restriction to 
pure empiricism” to mean that so long as one is at the experimenting 
table one investigates purely empirically, but for the interpretation of 
the results obtained uses “common sense,” that is, the traditional phi- 
losophy. It is these physicists who are most active in the movement 
against the world picture of the school philosophy. They are the physi- 
cists who try in the entire realm of their world conception to admit 
as an element only that which has been concretely experienced, as 
every physicist does at the experimenting table. 

These critically thinking physicists must ask themselves: how are 
those problems constituted the solutions of which represent advance- 
ment, as compared with the problems in the investigation of which 
scientists have been, figuratively, rotating about their own axes for 
centuries? 


97 



modern science and its philosophy 


For example, in the past the identity of light and electromagnetic 
radiation was not known, and now it is known. What does this sig- 
nify? By means of electrical devices— for example, radio transmitters 
—and by means of light sources one can produce phenomena that obey 
the same formal laws, the wave laws, and are differentiated by only 
one quantity, the wavelength. This recognition of the identity of light 
and electromagnetic radiation can be expressed as a perfectly definite 
statement concerning concrete experiences. It is not at all necessary 
to express it in such a way that something is said about the “nature” of 
light or electricity. Definite rules assign to the electromagnetic experi- 
ences symbols, the field quantities, among which there exist formal 
relations, the field equations. From given combinations of symbols 
one can then, using mathematical operations, derive new combina- 
tions with the help of the equations. These combinations may be trans- 
lated into experiences again with the help of the same rules of assign- 
ment. Hence, with the help of the theory, which consists of rules of 
assignment and of field equations, one can draw conclusions about future 
or past experiences on the basis of given experiences, In this way one 
has practical control over the experiences. Identity of light ind elec- 
tricity then means an identity of mathematical relations between sym- 
bols. From the theoretical point of view, the solution of a problem 
means the assignment of symbols to experiences, among which there 
exist relations that can be stated. From the more practical point of 
view, it means the possibility of obtaining control over one’s experi- 
✓ences with the help of this system of relations. 

We see then that in no problem of this sort is it ever a question 
of bringing about an “agreement between thought and object,” as the 
school philosophy says. Rather, it is always only a question of inventing 
a procedure which, with the help of a skillfully chosen system of sym- 
bols, is capable of bringing order into our experiences, thus making it 


easier for us to control them. Truth cannot be sought nf our 

experiences. The aim of the investigation is not the seeking after a 


“l yality. hidden in a nnts>|c ll ” The edifice of science must be built up 
</out of our experiences and out of them only. 


Before I go further into a discussion of how much the present physi- 


98 



^Iwentieth-century physics and school philosophy 

cal theories require such a conception of science, I should like to show 
with the help of a few historical remarks how the edifice of the school 
philosophy was gr adually underm ined, and what new conceptions have 
taken its place. As this development is still in its beginning stage, I shall 
have to give a somewhat aphoristic presentation rather than a sys- 
tematic one. 

In the city of Prague there lived and wrote the physicist who led 
most determinedly the struggle against the conceptions in physics 
that correspond to the reality concept of the school philosophy. Erns t 
Mach t aught in Prague from 1867 to 1895, that is, from his twenty- 
ninth to his fifty-seventh year. He was professor of experimental 
physics, first at the then bilingual University of Prague, and after the 
separation, at the German University. Here he wrote those of his 
works that are most important f or the epistemology of physi cs; “The 
History and Root of th e Principle of Cons ervation of Energy” (1871),* 
and “Mechanics in its Development” (1883).* 

His fundamental point of view was that all principles of physics 
are principles concerning the relations between sense perceptions, 
hence principles that state something a bout concrete experien ces. All 
concepts such as atom, energy. forc B, and matter are, according to 
Mach, only auxiliary concepts, allowing one to make statements about 
sens e perceptions in a simpler and more syn ogti Ejorm. tban if they were 
formulated directly as statements about the perceptions. In this way, 
all questions concerning the nature of force, matter, and so on, be- 
come meaningless, for these concepts can be eliminated from all physi- 
cal statements so that only statements about concr ete experiences ar e 
left. The ignorabimus toward the question of the n ature of matter and 
force, according to this conception, has no more justification than if a 
mathematician were to say: "Science, to be sure, can set up aU the 
theorems about complex numbers, but it will never be able to explain 

^Die Oeschichie tmd die Wurzel des Satzes von der Erhaltung der Arbeit, 
tr. by Philip E. B. Jourdain as History and Root of the Principle of the Conservation 
of Energy (Chicago: Open Court I^hlishing Co., 1911). 

’ Die Mechanik in threr Entioickelung, historisch-kritisch dargestellt, tr. hy 
Thomas J. McCormack as The Science of Mechanics (Chicago: Open Court Pub- 
lishing Co., 1803). 


99 



modern science and its philosophy 


i 


For example, in the past the identity of light and electromagnetic 
radiation was not known, and now it is known. What does this sig- 
nify? By means of electrical devices— for example, radio transmitters 
—and by means of light sources one can produce phenomena that obey 
the same formal laws, the wave laws, and are differentiated by only 
one quantity, the wavelength. This recognition of the identity of light 
and electromagnetic radiation can be expressed as a perfectly definite 
statement concerning concrete experiences. It is not at all necessary 
to express it in such a way that something is said about the “nature” of 
light or electricity. Definite rules assign to the electromagnetic experi- 
ences symbols, the field quantities, among which there exist formal 
relations, the field equations. From given combinations of symbols 
one can then, using mathematical operations, derive new combina- 
tions with the help of the equations. These combinations may be trans- 
lated into experiences again with the help of the same rules of assign- 
ment. Hence, with the help of the theory, which consists of rules of 
assignment and of field equations, one can draw conclusions about future 
or past experiences on the basis of given experiences. In this way one 
has practical control over the experiences. Identity of light ind elec- 
tricity then means an identity of mathematical relations between sym- 
bols. From the theoretical point of view, the solution of a problem 
means the assignment of symbols to experiences, among which there 
exist relations that can be stated. From the more practical point of 
view, it means the possibility of obtaining control over one’s experi- 
✓^nces with the help of this system of relations. 

We see then that in no problem of this sort is it ever a question 
of bringing about an “agreement between thought and object,” as the 
school philosophy says. Rather, it is always only a question of inventing 
a procedure which, with the help of a skillfully chosen system of sym- 
bols, is capable of bringing order into our experiences, thus making it 
easier for us to control them. Truth cannot be sou ght nntsidp-o f our 
experiences. The aim of the investigation is not the seeking after a 
“reality, hidden in g niifubpll ” The edifice of science must be built up 
»/out of our experiences and out of them only. 

Before I go further into a discussion of how much the present physi- 

98 



^Iwentieth-century physics and school philosophy 

cal theories require such a conception of science, I should like to show 
with the help of a few historical remarks how the edifice of the school 
philosophy was gr adually underm ined, and what new conceptions have 
taken its place. As this development is still in its beginning stage, I shall 
have to give a somewhat aphoristic presentation rather than a sys- 
tematic one. 

In the city of Prague there lived and wrote the physicist who led 
most determinedly the struggle against the conceptions in physics 
that correspond to the reality concept of the school philosophy. Erns t 
Mach t aught in Prague from 1867 to 1895, that is, from his twenty- 
ninth to his fifty-seventh year. He was professor of experimental 
physics, first at the then bilingual University of Prague, and after the 
separation, at the German University. Here he wrote those of his 
works that are most important f or the epistemolo g y of physi cs: “The 
History and Root of the Principle of Conservation of Energy” (1871),® 
and “Mechanics in its Development” (1883).® 

His fundamental point of view was that all principles of physics 
are principles concerning the relations between sense perceptions, 
hence principles that state something a bout concrete experien ces. All 
concepts such as atom, energy, force, and ma H-er are, according to 
Mach, only auxiliary concepts, allowing one to make statements about 
se nse perceptions in a simpler and more syno pt ic J ornu than if they were / 
formulated directly as statements about the perceptions. In this way, 
all questions concerning the nature of force, matter, and so on, be- 
come meaningless, for these concepts can be eliminated from all physi- 
cal statements so that only statements about concrete experiences ar e 
left. The ignorabimus toward the question of the natu re of matter and 
force, according to this conception, has no more justification than if a 
mathematician were to say: “Science, to be sure, can set up all the 
theorems about complex numbers, but it will never be able to explain 

Geschichte und die Wurzel des Seizes von der Erhaltung der Arbeit, 
tr. by Philip E. B. Jourdain as History and Root of the Principle of the Conservation 
of Energy (Chicago; Open Court Publishing Co., 1911). 

® Die Mechanik in ihrer Entwickelung, historisch-kritisch dargesteUt, tr. by 
Thomas J. McCormack as The Science of Mechanics ( Chicago; Open Court Pub- 
lishing Co., 1893). 


99 



modern science and its philosophy 


the nature of the complex number. Toward this problem we must 
modestly acknowledge an eternal ignurabimus." To this contention 
every other mathematician will respond that the complex numbers 
have been introduced only in order to make clearer certain statements 
about real numbers, that basically every theorem from the theory of 
functions of a complex variable can be expressed also as a theorem 
about real numbers. 

Neither Mach himself nor his immediate students has systematically 
carried further his point of view and set up, in opposition to the world 
conception of the school philosophy, a similarly coherent scientific con- 
ception. On the contrary, Mach’s teaching, through many presentations, 
has been washed out into something indefinite rather than built up to a 
consistent scientific conception of the world. It has even been inter- 
preted again in line with the school, philosophy, sometimes more re- 
alistically, sometimes more idealistically, so that to many— as, for ex- 
ample, to the great anti-Machi stic school in Russia, at the head of 
which stood Lenin himself— it appeared not as the beginning of a new 
scientific world conception but as a new fashionable form o f the school 
philosophy. 

Conceptions similar to those of Mach were presented in France, in 
part independently, by the physicist Pi erre Duh em. In his expositions, 
Qubem did not equal Mach in breadth of vision, but often surpassed 
him in sharpnes^_gfJogic. 

From an entirely different direction there arose against the school 
philosophy a movement often referred t o as “conventionali sm." Its most 
important representative is the French mathematician, physicist, and 
astronomer, Henri Poincar e. He called attention to the fact that physi- 
cal principles often contain concepts that are defined by these very 
principles. In such cases the principles can never be tested against ex- 
perience, since they are disguised definitions, “conventions.” Thus, ac- 
cording to Poincare, the law of conservation of energy is nothing but a 
definition of the concept of energy . The significance of conventionalism 
for the understanding of what is expressed by the principles of physics 
is very great, in my opinion, and perhaps no one among the physicists 
has contributed as much as Poincare to the shaking of the school phi- 

100 



^twentieth-century physics and school philosophy 

losophy. In Germany the chief representative of this movement is Hugo 
Dingier. Through extreme application of conventionalism, however, 
Dingier has again approached the school philosophy, according to the 
principle that opposites meet, by attempting to show that certain con- 
ventions are the simplest and hence are the only ones justified. 

A direct attack against the truth concept of the school philosophy 
was made by the American psychologist William James in his boo k 
Pragmatis m, which introduced the pragmatic movement that has spread 
so vwdely. According to James, the truth of a system of principles— a 
physical theory, for instance— does not consist in its being a fa ithful copy 
of realit y, but rather in its allo wing us with th e help of_&es e principle s 
to change our experiences according to our wishes. According to this 
view, which essentially agrees wiA that of Mach, but rejects even 
more bluntly the truth concept of the school philosophy, every solution 
of a problem is the constructionof a proced ure that can be of use to us in 
the ordering and mastering of our experiences. If, for example, we are 
familiar with all the means and rules of machine construction, if we 
know what motion takes place under given conditions, then it is clear 
that it would not help us further in the slightest if we knew, besides this, 
the nature of matter and force. If we understand the solution of a prob- 
lem in the sense of James, then we cannot, in general, consider ques- 
tions like the latter as scientific formulations of problems. 

Henry Bergson, in his introduction to the French translation of 
James’s Pragmatism, from which we have taken the characterization of 
school philosophy, also characterizes very clearly and pertinently the 
contrasting pr agm atic conce pHnn nf truth and 

The other conceptions [those of the school philosophy] represent truth 
as something that was present before the w ell-defined act of the man who 
formulated it for the first time. We say he was the first W see it, but it had 
been waiting for him as America had waited for Columbus. Something had 
hidden it until that moment from all eyes, had covered it, so to speak, and 
he had discovered it. Quite different is the conception of William James. 
He does not deny t hat reality ^ at least to a large extent, is independent of 
what we say or think. But truth, which can only be associated with what we 
state about reality, appears to him to be created by our statement. We in- 
vent truth in order to make reality useful to us, just as we create mechanical 


101 



modern science and its philosophy 


devices in order to make the forces of nature useful to us. It seems to me 
that one might summarize the essence of the pragmatic conception of truth 
in a formula of the following kind: Whereas for th e other conceptions a 
new truth is a discovery, fnr pragmatism it is nn invpntinn 

The objection has often been raised that pragmatism correctly char- 
acterizes only the practical, not the theor etical, significance of science. 
To this objection James himself replied that next to the interest that man 
has in breathing freely his greatest interest, which in contrast to most 
interests of a purely physical kind knows no fluctuation and no failure, 
is the interest in feeling that he is not contradicting himself, that what he 
thinks at the present moment agrees with what he thinks on other oc- 
casions. As we shall soon see, however, unambiguity or freedom from 
contradictions is the most essential attribute of every cognition; hence, 
confiicts between the practical and theoretical conceptions of truth do 
not arise from the doctrine of pragmatism. 

The physicist in his own scientific activity has never employed any 
other concept of truth than that of pragmatism. The “agreement of 
thoughts with their object,” as the school philosophy requires, cannot be 
established by any concrete experiment. In practice we encounter only 
experiences, never an object; hence nothing can be compared with an 
object. Actually, the physicist compares only experiences with other ex- 
periences. He tests the truth of a theory through what one is accus- 
tomed to call “agreements.” 

Thus, for example, one always obtains the same numerical value 
for the Pl anck constant h . using various methods. This really means that 
the quantity h can be constructed in entirely different ways from experi- 
ences— as from the experiences of black-body radiation, or those of the 
Balmer series of the hydrogen spectrum. The theory in which h plays a 
role then asserts that all the various groups of experiences, which are 
qualitatively so different from one another, nevertheless should give the 
same numerical value of h. It is therefore only a question of comparing 
experiences with one another. This procedure, which the physicist is ac- 
customed to use in his work, has been made by Mach and James into a 
general conception of the criteria of trutji. 

In all this it must be admitted, however, that these conceptions pos- 

102 



twentieth-century physics and school philosophy 

sess a certain indefiniteness for the mathematical physicist. He always 
has the impression that there is a lack of precision. In particular, he 
finds it hard to take the pragmatic theory of truth quite seriously. This 
arises partly from the fact that James, and to a certain extent Mach also, 
failed to set a very high val ue on the role of formal logic in the cnnst mc- 
tion of the system of human cognition. In fact, in a certain opposition 
to the misuse of logic by school philosophy, they both emphasized the 
fluid elements in cognition as against the rigidly logical. Another way 
of putting it is that, in opposition to the viewpoint of mathematica l 
logic, which to them always smacked of the school philosophy, they 
presented the viewpoint of evolutionarY biolo gy. Because of this the 
mathematicians and mathematically thinking physicists have often been 
forced into a certain opposition to the doctrines of Mach and James. 
Many of them, attracted by the logical garments of the school philos- 
ophy, have even behaved in a friendlier way toward the latter than 
toward the modem tendencies. 

It is therefore significant that the school philosophy has also been 
criticized from quite another side; indeed, that position was assailed 
which seemed most unassailable, namely, the logic of the school phi- 
losophy. The logic employed by philosophers until well into the nine- 
teentn century was not very different from that formulated bv Aris - 
totle. In connection with the investigations on the foundations of 
mathematics, however, there developed in the field of logic a fresh 
tendency, which shook the old s cheme of Aristo tle. This movement is 
represented in Germany, especially, by the names of Schroder, Freg^ 
and Hilbert. Through the use of a symbolism modeled afrp.r that nf 
mathematics it gave to logic ^ flex ibility_and_ free dom of ino vement 
that it had not possessed before. This made it possible to deal with far 
more complicated thought structures than could have been handled 
on the basis of the school logic. 

It turns out, in fact— and this is chiefly the work of the English , 
mathematician and logician Bertrand Russell and his students, par- 
ticularly the Austrian Wittgenstein— that the logic of the school philos- 
ophy, because of the narrowness of its sc heme, made it impossible 
from the very start to express certain thoughts. Thus, many of the 


103 



modern science and its philosophy 


principles regarded by the school philosophy as being certain were 
certain only because the contrary did not fit into the scheme of Aris- 
totelian logic. 

In this way Russell pointed out that one of the most fateful er- 
rors of the school logic was its assumption that every judgmen t con- 
sists in attributin g^ to a sub j^^*^ pmpprty as_prfidicat:e. If one says, 
for example, that a body A is moving with respect to another body B, 
the adherent of the school logic will demand that to one or the other 
of the two bodies shall belong the predicate of motion. Now Russell 
has shown that very many judgm ents-consist i n stating a r elat ion be- 
twee n twn things a nd cannot be r educed to the statement of a property 
of a single thing, which is a much more special case. However, to the 
adherent of the s^ool logic a statement like the following, for ex- 
ample, appears nonsensical: “If two bodies are moving with respect 
to each other, it is meaningless to ask which one is ‘really’ moving, that 
is, to which one belongs the predicate ‘being in a state of motion.’ ” 

With regard to the school philosophy, which has taken over the old 
logic more or less consciously, Russell remarks that it is led by the 
unconscious conviction that all statements of judgments must have the 
subject-predicate fo rm— in other words, that every statement must at- 
tribut e a property ta .a_thing. This conviction has made most philoso- 
phers incapable of understanding the wo rld of science and of everyday 
life. According to Russell, however, most philosophers are less con- 
cerned with attaining ^ true under standin g tlii&- w orld than with 
showing its unreality in the interest of a transcendental, truly real 
world. 

With the aid of the old logic the school philosophy could easily 
deduce the absurdity of the pragmatic concept of truth and the rela- 
tivistic conception of physics. On the other hand, the ne y logic of Rus- 
sell and his school was suitable to help build up th e pmely empiri cal, 
and hence still somewhat vague, conceptions of Mach and James into a 
real system of the scientific world conception that was superior to the 
school philosophy from the standpoint of f ormal logic as well. 

’There were some philosophers with a mathematical-physical orien- 
tation who followed Russell but at the beginning cared little for Mach 



twentieth-century physics and school philosophy 


and almost nothing for James. Nevertheless, like the latter, they too 
discarded the truth concept of the school philos ophy. In contrast to 
the method of jj^n ^tism . however, they not only tried to characterize 
the system of science in a general and somewhat indefinite way by 
saying that the system is an instrument to be invented and constructed 
in order to find one’s way among experiences, but also— and instead— 
they investigated the structure of this instrument. The investigation 
took place through an analysis of the method by which physics orders 
the experiences through a mathematical system of formulas. From this 
most advanced science, as an example, one can get an insight into the 
requirements that must be imposed on scientific cognition in general. 

What then are the elements of which the instrument that we call 
science or cognition is composed? Here the infiuence of the mathe- 
matical-logical movement becomes felt. The new epistemology says 
that the system of sciencb'^nsists of symbols. This conception has 
been most clearly formulated by Moritz Schlick i n his “General Theory 
of Knowledge.” * Lik e Jame s, Schlick begins with a determined rejec- 
tion of the jruth concept of school philosophy. He says that 


in the past, the truth concept was almost always defined as the agreement of 
the thou ght wit h its objects. 

« 

He then shows that the word agreement” cannot mean here anything 
like equality or similarity, as in ordinary usage, since between a judg- 
ment and the circumstances that it judges there can be no similarity. 
He continues: 


Thus the concept of agreement melts away before the rays of analysis, 
in so far as it is to mean equality or similarity; and what is left is only a 
unique correspondence. In diis consists the relation of the true judgment 
to reality, and all those naive theories according to which our judgments 
and concepts somehow could portray reality are completely destroyed. 
There remains for the word "agreement” no other meaning than that of 
unique correspondence. One must dismiss the thought that a judgment in 
relation to the facts of the case could be anything more than a symbol, that 
it could be more intimately connected with them than through mere cor- 
respondence, that it is able somehow to describe or express or portray them 

*M. Schlick, Allgemeine ETkenntnislekre (Berlin: Springer, ed. 2, 1925). 

105 



modern science and its philosophy 


adequately. Nothing like it is the case. The judgment portrays the nature 
of what is judged as little as the note the tune, or as tlie name of a man 
his personality. 

If a person had always known and kept in sight the fact that cognition 
arises simply through an assignment of symbol to object it would never have 
occurred to him to ask whedier it is possible to have a cognition of things 
“as they really are” themselves. He would be led to this problem only by 
the opinion that cognition is a kind of pictorial representation that portrays 
the things in one’s consciousness. Only under this assumption could he ask 
whether the images really have the same qualities as the things themselves. 

It is easy to convince oneself that physical cognition consists in 
the unequivocal assignment of a system of symbols to experiences. 
Thus, for example, to electromagnetic phenomena are assigned field 
intensities, charge densities, and material constants as symbols. Among 
these symbols there exist formal mathematical relations, the field equa- 
tions. Symbols that are equivalent to one another according to these 
relations or to the general logical and mathematical laws can be as- 
signed to the same experiences without violating the uniqueness in the 
required sense. 

If, for example, we start with a definite measured distribution of 
electric charge on the surface of a sphere, there is assigned to this ex- 
perience, as a symbol, a definite mathematical function— the charge 
density as a function of position. If the sphere is left alone and we ex- 
amine it after a lapse of some time, we measure everywhere the same 
charge density. To this experience is assigned a constant number for 
the density. If now the field equations were so constituted that accord- 
ing to them there would be obtained for the calculated charge density 
after some time a function other than a constant, then we should have a 
symbol system that would assign to the final electrical state of the 
sphere various symbols that are not equivalent. Because of this am- 
biguity we should say that our symbol system, which on the one hand 
is based on the rules of assigning symbols (here it is the method of 
measuring electric charges), and on the other hand on the relations 
among the symbols (here the field equations), gives no true cognition 
of the electrical phenomena. 

Every verification of a physical theory consists in the test of whether 



^entieth-century physics and school philosophy 

the symbols assigned by the theory to the experiences are unique. If, 
for example, the Planck constant h occurs in the equations, it denotes 
a definite experience. This can be produced concretely if by means of 
the equations we express h through so-called “observable” quantities, 
that is to say, those symbols to which our rules assign concrete experi- 
ences. In this way an experience is then directly assigned to the quan- 
tity h. As is well known, one can express h through quantities con- 
nected with the observation of black-body radiation and through 
quantities that arise from the observation of the Balmer series in the 
hydrogen spectrum. There are thus two experiences apparently de- 
noted by h. These consist in the calculation of its value from two dif- 
ferent groups of phenomena. If the two gave different values for h, 
we should be denoting two entirely different experiences by the same 
symbol h. We should then have in the system of equations involving 
h, in conjunction with the rules of assignment (methods of measure- 
ment), a symbol system that does not denote the experiences uniquely, 
and hence represents no true cognition. Through the fact that they both 
do give the same value of h we recognize the uniqueness of the sys- 
tem, of symbols, the “truth” of the theory. 

This comparison of the values of a quantity, calculated in different 
ways from observations, is the only way in which the physicist in his 
actual work can test the “truth” of a theory. The direct comparison of 
“observed” and “calculated” values, as it is often called in works on 
physics, turns out on closer examination to be nothing else than the 
test of the uniqueness of a symbol system. Suppose, for example, that 
on the one hand I calculate a current strength from the electronic the- 
ory of metals, while on the other hand I “observe” a galvanometer. 
Then this alleged observation is also only a calculation from another 
theory, namely, that of the galvanometer. For in reality I observe only 
the coincidences of pointers and scale divisions, and even these sub- 
stantiations on more careful analysis would turn out to be results of a 
theory of the solid body. Even in the limiting case, where a value “is 
observed as directly as possible”— for example, if it is a question only 
of the position of a pointer on a scale— it is still calculated from the 
theories of rigid bodies and light rays, because I observe directly only 


107 



modern science and its philosophy 


dancing spots of color and not positions of the pointer. Hence what I 
usually call the comparison of observed and calculated values is, for 
example in our case, the comparison of the values of currents given 
from two different theories for the same concrete experience. 

The school philosophy interprets an agreement of this kind— 
which, we are convinced, is the only criterion of truth for tiie physicist 
—in the following way: If for a quantity, for example, h, the same 
numerical value is obtained in various ways, then this quantity has a 
real existence. If by this expression is understood only what was 
really substantiated, namely, that the quantity h occurring in the equa- 
tions can be calculated uniquely from phenomena of various kinds, 
then no objection can be raised against it. However, we must not say 
then that from the agreement of the results of the measurements we can 
draw conclusions about the real existence of h; for in that case this 
existence is identical with agreement. 

Similarly, no conclusion drawn from physical experience concerning 
the real existence of the quantum of action, the elementary charge of 
electricity, or similar concepts, is a scientific conclusion. It finds its 
justification only in the metaphysical representation of reality given 
by the school philosophy, according to which the true principles exist 
before all experience and must be discovered by investigation, as Co- 
lumbus discovered America. 

1 believe that the mathematician can get from the following il- 
lustration a very good understanding of the difference between the 
school philosophy, which recog nizes metaphysical reali^T and the sci- 
entific world concepti on, which uses only constructions based on con- 
crete experiences. 

If I have a convergent sequence of rational numbers, having an 
irrational number as a limit, I can establish a convergence without 
making use of the concept of the irrational number. That is to say, I 
need only to establish that the difference of any two rational members 
of the sequence above a given index can be made arbitrarily small by 
choosing an index that is suflBciendy large> I have therefore before me, 
if I have defined only the concept of the rational number, a sequence 
of rational numbers having the property of convergence, but no limit 


108 



twentieth-century physics and school phiiosophy 

in the domain of rational numbers. As is clear to every mathematician, 
there is no conclusion by the help of which it could then be shovm 
that a limit of this convergent sequence exists. Rather, the convergent 
sequence itself is the concrete exhibitable object. However, I can now 
define such a sequence as an irrational number. This means that in all 
theorems involving irrational numbers I can substitute for the latter, 
sequences of rational numbers. It is not necessary, and is not justifiable 
logically, to speak of a real existence of irrational numbers, inde- 
pendent of the rational ones. 

If we take this as an analogy, then the concrete experiences cor- 
respond to the rational numbers and the so-called really existing truths 
to the irrational numbers. A group of experiences with a system of 
symbols assigned to it, in which such agreements can be established 
throughout as we established in the case of the constant h, for example, 
corresponds to a convergent sequence of rational numbers. The 
uniqueness of the symbol system can be established within the group 
of experiences itself vidthout having recourse to an objective reality 
situated outside, just as the convergence of a sequence can be estab- 
lished without the need of discussing the limit itself. 

Likewise, as the concept of the irrational number is first defined by 
the convergent sequence of rational numbers, so the concept of true 
existence, say of the quantum of action h, is first defined by the agree- 
ments in the whole group of experiences involving h. 

Just as the expression “irrational number” is an abbreviation for a 
convergent sequence of rational numbers, so the concept of a really 
existing quantum of action is only an abbreviation for the group 
of experiences which yield one and the same numerical value 
for h. 

It is completely false to say— as is often said— that the agreements 
of the values of h are explained most naturally by means of the hypothe- 
sis of the real existence of a quantum of action. In a hypothesis one can 
only state a conjecture about future experiences, not about the “real 
existence” of a thing corresponding to an assigned nam^To state a 
conjecture of this kind would be exactly as if a mathematician were to 
say: “The existence of convergent sequences of rational numbers with- 


109 



modem science and its philosophy 


out limits can be explained most naturally by means of the hypothesis 
that there are irrational numbers.” In reality, through such an asser- 
tion he would only be giving a new name to convergent sequences 
without limits. Similarly, through the assertion of the existence of a 
quantum of action no new fact is stated besides the agreements. Hence, 
also, no hypothesis is present, 

I have already spoken about the fact that the development of the 
scientific conception of nature was retarded by a certain conflict be- 
tween the mathematical-logical and the biological-pragmatic approach. 
The latter was greatly lacking in precision, so that even the school 
philosophy appeared to have many advantages here. Thus, for ex- 
ample, Bertrand Russell in his book. Our Knowledge of the External 
World‘ in many points showed greater concurrence witli the concep- 
tions of the school philosophy than with those of Ernst Mach. In the 
German translation of this book, however, Russell remarks in a foot- 
note that he now agrees with Mach on one of the most important 
points. It seems to me that in the case of other representatives of the 
Russell movement as well the conviction is gaining ground that a con- 
sistent further expansion of the scientific world picture is not to be 
sought by opposing the Mach conception in favor of the school phi- 
losophy for the sake of its apparently more rigorous logic. On the con- 
trary, it appears that this expansion must be carried out with the aid 
of modem logic by building up the doctrines of Mach to form a sys- 
teip everywhere free from objections on logical grounds. 

Although according to modem conceptions logic can produce 
nothing more than tautological tran.sfo rmations of principles, it is 
nevertheless indispensable for the construction of a rigorously scientific 
world picture. The reason is that many of the prejudices of the school 
philosophy arose from the fact that me re tautologie s were regarded 
as expressions of cognition. A complete survey of all possible tautologi- 
cal transformations therefore offers the possibility of constmcting on 
the basis of Mach’s views a scientific edifice that is superior to the 
school philosophy in its logical precisiom 

'The most determined attempt in this direction was undertaken by 
' Chicago: Open Cou^ Publishing Co., 1914. 


110 



twfentleth-century physics and school philosophy 

Rudolf Carnap. In his book “The Logical Structure of the World,”* 
which appeared in 1928, he seeks to build up the whole system of 
science, starting from concrete experiences. He tries to show that all 
principles in which p hysical or psychologica l objects are inyolyed can 
be replaced by statements concernin g concrete experi ences. The rules 
according to which statements about concepts must be replaced by 
statements about concrete experiences are called by Carnap the con- 
stitution of these concepts. In a scientific statement there should occur 
only concepts the constitution of which is known. The basis of eyery 
science is the system of concept constitutions. The building of this 
system step by step with the help of the modem lope o f Russell is*^ 
w^t Camap calls the logical structure of the world. 

' According to this conception, a scientific problem can consist only 
^in asking whether a definite scientific statement is tme or fals e. Since, 
howeyer, eyery such statement can be reduced to a statement about 
concrete experiences if the constitution of the concepts occurring in 
it is known, eyeiy problem that can be called scientific consists in the 
question whether a definite relation among concrete experiences does, 
or does not, exist. At the same time Camap shows further that all rela- 
tions, in the last analysis, can be reduced to statements of a similarity 
between concrete experiences. Since one can properly assume that 
eyery such similarity is confirmable in principle, it follows that eyery 
problem that can be scientifically formulated is also soluble in prin- ^ 
ciple. ' 

We see that the consistent carrying through of a piurely scientific 
world conception, as attempted by Camap, leads us just as far away 
from the resigned ignorabimus as does the pragmatism of James, 
which is thought out somewhat less logically but in its tendencies has 
the same goal. Camap formulates it thus: 

Science, the system of conceptual co^^on, has no limitations . . . 
There are no questions the answers to which are impossible for science in 
principle . . . Science has no boundary points . . . Eyery statement 
formed ^from scientific concepts can be established in principle as tme 
or false. , 

*R. Camap, Der logische Aufbau det Welt (Berlin, 1928). 

Ill 



modern science and its philosophy , 

r 

( This does not mean, of course, that there are no other domains of 
life than science. These domains, however— lyric poetry, for example- 
are distinct from science. Through the latter itself no problems can be 
set up that are insoluble by its means. 

Wittgenstein says very precisely; 

an answer cannot be expressed, neither can the question be 

expressed. 

Hence, as understood by Carnap and Wittgenstein, the questions 
beloved by the school philosophy, such as whether the outer world 
really exists, not only cannot be answered but cannot even be expressed, 
because neither the positive assertion, falsely called the realistic 
“hypothesis,” nor the negative idealistic assertion can be expressed 
through constituted concepts. In other words, neither assert ion can be , 
expressed as a substantiable relation among concrete experiences. We 
see here the close relationship betwe n the truth concept of the m odem 
l ogical movement and that of prag matism. 

A tendency similar to that of Schlick and Carnap is followed by 
Hans Reichenbachj but Reichenbach departs from Carnap in many 
points, as in the recognition of the realistic stan dpoint. 

After this survey of the movements that seek to construct a purely 
s cientific world conception through following closely the actual prac- 
✓• ^ce of mathe matical and physical investigati on in contrast to the school 
philosophy, let us return to our starting point, the question of why 
physicists often refuse to pass judgment on problems like spa ce, time, 
and causality, and leave them to the philosophers. 

This refusal, we can now say, arises from the fact that these physi- 
cists, consciously or unconsciously, cling to the doctrines of the school 
philosophy, according to which such problems must be solved by 
methods fundamentally different from those employed by physicists. 

If a scientist follows these thoughts through logically, he must end up 
in the blind alley of ignorabimusT^ 

If, however, we stand on the ground of the purely scientific world 
conception, we know that the solution of a scientific problem can only 
consist in finding out new relatio ns among concrete exp eriences, or. 


112 



twentieth-century physics and school philosophy 


to express it in another way, in making progress in the unique designa- 
tion of experiences by a system of symbols. 

One can try to fit new experiences into places in the existing sys- 
tem of symbols; this we call purely experimental investigation. The 
idea that there can be a still pmer type of experimental investigation 
that makes no use at all of symbol systems is, in my opinion, an illu- 
sion.^To be sure, as Schlick properly points out, one can experience 
phenomena and can get to know them without using any symbol sys- 
tem; but this is not scientific cognition. For, at best, one could then 
establish, for example, that today about noon several colored spots were 
seen in certain combinations, although a more exact analysis would 
probably show that even in such an utterance there is already an as- 
signment of symbol to experience. 

The work of the theoretical physicist consists, in part, in investigat- 
ing the consequences resulting from the fundamental relations be- 
longing to the symbol system. This is an essentially mathematical task, 
such as, for example, the integration of the field equations, the funda- 
mental relations among the field quantities. Another part of the work 
of the theoretical physicist consists in extending the symbol system. 
Naturally, with the introduction of new symbols new rules of assign- 
ing them to experiences must also be introduced. 

If, for example, in the investigation of the hardness of a substance 
one must make a new hypothesis concerning its crystal lattice, this 
means a change in the symbol system; specifically, in the geometric 
figure by which the substance concerned is characterized. Everyone 
will acknowledge that a work of this kind is really a concrete physical 
work. From such changes in the symbol system there extends a con- 
tinuous series to those changes that the physicist often feels to be 
“speculative” or “philosophical,” as, for example, the introduction of 
the Einstein time scale. Here, too, the question is only one of setting 
up a new rule of assigning the symbols t and If in our equations to 
our experiences, as well as of a new relation between the symbols 
f and If in the symbol system. One can give no criterion, however, to 
determine whether a change means a physical cognition or a philo- 
sophical one. For the physicist there exist no such limitations. Whether 


113 



modern science and its philosophy 


I am dealing with the measurement of hardness or with that of space 
and time, it is always only a question of assigning a unique system of 
symbols to experiences. Nowhere is there a point where the physicist 
must say: “Here ends my task, and from here on it is the task of the 
philosopher.” 

This could happen only for a thinker standing on the ground of the 
school philosophy. He might ask, for example: “When I have exhausted 
all problems of assigning time symbols to experiences, which among 
the time scales admitted by the relativity theory is the true, real time?” 
And that question the physicist cannot answer; it is the philosopher 
who must pass judgment on it. 

It looks as if classica l physics ha s been living on good terms with 
school philosophy, whereas modem physics, with its relativity theory 
a nd quantiiTn mechanic s- has at once come into conflict with it. The 
physicists who feared the break with school philosophy could resolve 
this conflict only by a kind of doctrine of double truth. They said, in 
effect; “We physicists speak on ly of time- measurements; hence for us 
the relativity theory is valid. The philosopher speaks of t he real 
time; for him perhaps something else is valid.” If this doctrine of 
d ouble truth wa s meant somewha t ironcially, as was often the case, 
it was an irony of embarrassment. There are actual cases, however, 
when it is meant in all seriousness. 

The reason that the conflict did not occur during the time of classical 
physics is simply that, for example, the time concept of the school 
philosophy had just as much an empirical physical origin as that of the 
relativity theory, the difference being that the former corresponded to 
the o lder state of physics, that which we today call the classical. The 
symbol system with the help of which Newtonian mechanics and 
Euclidean geometry portrayed space and time experiences was de- 
clared by the school philosophy to b e real s pgf^p anfl Hmf»^ and pro- 
claime d as eternal tru th. 

If, however, in accordance widi the scientific world picture, we 
regard every problem as merely one of assigning symbols to experi- 
ences, then in the designation of the space and time experiences a 
change is fully as possible as in the rest of physics. Just as this leads to 

114 



twpntieth-century physics and school philosophy 


progress in the theory of the solid body, so it also leads to progress in 
the study of space and time, keeping up with the progress of our 
observations. 

One cannot declare certain parts of the symbol system to be un- 
changeable for all time. To be sure, one may keep the old Kantian 
• terminology, in a certain sense, and explain space and time as the 
frames of physical phenomena. But one must bear in mind, as Reichen- 
bach correctly says, that this frame, too, must be always adapted more 
and more closely to our progressing expIfiSice. 

The rise of the physics nf Galileo and Newt on brought about the 
breakdown of the philosQoh v- of Aristo tle, which attempted to show the 
eternal truth of the physios of antiquity. Similarly, beside the relativity 
t heory and quantum^ meoh anics there cannot e.xist a philosophy that 
contains a fossiliza tion of jhe yi grlier phys ical theories. 

Just as the views of the school philosophy on space and time make 
it more difficult to understand the relativity theory, so its conception of 
causality is an obstacle to an understanding of the new quantum 
mechanics. I will not go into greater detail here concerning the causality 
problem; I wish to draw attention to only one point. 

Classical physics understood by the law of causality the calculahility 
of future states from an initial state: if the state of the world or of an 
isolated system is known exactly at one instant of time, then it is known 
also for all future time. It was regarded as indubitable that with the 
help of prescribable methods of measurement one could determine the 
values of the quantities defining the state of a system— if not exactly, 
then at least approximately. It was assumed that with increasing re- 
finement of the methods of measurement it would be possible to in- 
crease the accuracy arbitrarily. Hence, in principle, to such quantities 
as lengths or field intensities, determining the state, were ascribed exact 
numerical values. 

That people held this conviction so firmly is owing to the belief of 
the school philosophy that exact values of lengths or field intensities 
must exist even if they are not yet known accurately to the person per- 
forming the measurements. They are hidden in the nutshell referred to 
by Bergson, and one must break through it to reach the true values. 


115 



modern science and its philosophy 


( It is quite natural that we cannot measure the exact length of a 
rod. If I wish to assert, however, that through refinement of the 
methods of measurement we can gradually come closer and closer 
to the exact value of the length, it is first necessary to ask whether we 
can define what we understand by exact length. For here we often 
go in a circle. We define as exact value the limiting value approached 
by the measurements as the method is increasingly defined. In this 
definition we assume that such a limiting value exists. Its existence can 
be shown empirically, if at all, only to within an uncertainty of a 
definite order of magnitude. But in this way we have come no nearer 
settling the question of the existence of an exact value. 

According to the atomistic theory, the length of a rod is nothing but 
the distance between two atoms. Since an atom is a system of electrons, 
every such distance can be reduced to a distance between electrons. 
Every measurement consists in a comparison of the measured body 
with a measuring rod. But the latter is itself a system of electrons. 
Hence every measurement of length ultimately leads to the substantia- 
tion of a coincidence between electrons. By this, of course, is not meant 
a coincidence in the literal sense, but something like the phenomenon 
whereby one electron covers the other when viewed from a definite 
direction. It does mean, however, that the measurement of length is 
reduced to the observation of light diffracted or scattered by the two 
electrons. Now it is clear that differences in length that are small in 
comparison to the wavelength of the light involved can play no role 
in this observation. Such differences cannot be observed through any 
experiment of this sort, hence cannot be regarded as conceivable ex- 
perience. The possibility of being able to refine arbitrarily the measure- 
ment of length by refining the measuring technique can only be based 
on the hope of being able to produce radiation of arbitrarily small 
wavelength. The production of such radiation, which must therefore 
have arbitrarily high frequency, is not very probable, however, ac- 
cording to the observations that led to the formulation of the quantum 
hypothesis. For then there must be light quanta of arbitrarily high 
energy and exerting arbitrarily great forces on collision. Moreover, 
there is another circumstance, first pointed out by Heisenberg, that 



twentieth-century physics and school philosophy 

makes impossible an exact measurement even to within the order of 
magnitude of tlie errors that correspond to still attainable wave- 
lengths. If, specifically, one goes to very high frequencies, so that the 
collision force of the light quantum becomes very great, it is not pos- 
sible to establish the relative velocity— in our case, the state of rest 
relative to each other— of the two electrons, for the impulse imparted by 
the light quantum changes their state of motion in an uncontrollable 
way; this is the phenomenon that accounts for the Compton effect. 

Just as there is no method of measuring the length of a rod with 
arbitrary accuracy, so too there is no method of measuring the in- 
tensity of an electric field with arbitrary accuracy. Every such measure- 
ment is based on the observation of the force exerted on a test body 
in the field. The charge and size of this body are assumed to be so 
small that they do not disturb the field. This assumption, however, 
contradicts the atomistic hypothesis, which has no cognizance of arbi- 
trarily small and arbitrarily weakly charged test bodies. Therefore 
the assumption that a field intensity is measurable with arbitrary ac- 
curacy in principle is not justified. 

The physicist who starts from the conceptions of school philosophy 
must say to this argument that there really exist perfectly definite 
values of lengths, but that nature is so constituted that it prevents us 
from determining them. There are natural laws especially suited to this 
purpose. This corresponds entirely to the conception of the relativity 
theory according to which the absolute velocities of motion of aU 
reference systems exist but the laws of nature are so insidious that they 
prevent the observation of these velocities. The physicist who wishes to 
represent this conception of the relativity theory which corresponds 
to the school philosophy must assume the existence of realities to 
which no concrete experience corresponds. So, too, the physicist who 
assumes the existence of exact lengths of bodies must understand by 
the word “existence” something that no longer has any connection 
with the empirical sense of this word, which refers to something ex- 
perienced or, at least, experiencible. 

On the basis of this conception, the problem is then put: Is the 
law of causality valid or not valid in nature? That is, do the initial 


117 



modern science and its philosophy 


positions and velocities of electrons determine these quantities for all 
future time? Even if there were equations for wliich that were the 
case, nothing at all is said about real experiences. For we know that 
we cannot assign positions and velocities of electrons unequivocally to 
our experiences even by successive approximations. That the law of 
causality is not valid for our experiences involving the positions and 
velocities of electrons has been made plausible by the experiments on 
the diffraction of electrons as they are usually interpreted. If, to wit, 
electrons fall on a grating, the direction in which a single one is de- 
flected cannot be predicted from its initial position and velocity. 

It is often concluded that electrons follow absolute chance in their 
choice of direction, or even, as one occasionally finds it put in popular 
presentations, that an irrational element plays a role, a kind of per- 
sonification of the electron. This follows, however, only if one starts 
from the picture given by school philosophy, according to which every 
electron has a definite position and velocity, which nevertheless do not 
determine its future. 

From the standpoint of a purely scientific conception, on the other 
hand, one will say that there are no individual experiences involving 
positions and velocities of electrons from which the future of the latter 
can be predicted unequivocally. Instead, it appears that the probability 
that an electron will be deflected in a definite direction can be pre- 
dicted from the experience of the initial experimental arrangement. 
For these probabilities (the squares of the absolute values of the wave 
functions ) Schrodinger, in his wave mechanics, sets up rigorous causal 
laws. To the probabilities that occur in these laws and define the 
state of the system one can therefore assign definite experiences. This 
theory is called statistical. The statistical element here consists in the 
maimer of assignment of experiences to symbols. Thus to certain 
symbols, the squares of the absolute values of the wave functions, there 
are assigned, not individual experiences, but numbers which are ob- 
tained by averaging from a great many iiidividual experiences. 

The task of physics is only to find symbols among which there exist 
rigorously valid relations, and which can be assigned uniquely to our 
experiences. This correspondence between experiences and symbols 

I 

118 



twentieth-century physics and school philosophy 


may be more or less detailed. If the symbols conform to the ex- 
periences in a very detailed manner we speak of causal laws; if the 
correspondence is of a broader sort we call the laws statistical. I do not 
believe that a more exact analysis will establish a definite distinction 
here. We know today that with the help of positions and velocities we 
cannot set up any causal laws for single electrons. This does not ex- 
clude the possibility, however, that we shall perhaps some day find a 
set of quantities with the help of which it will be possible to describe 
the behavior of these particles in greater detail than by means of the 
wave function, the probabilities. When we determine a number 
through a so-called single observation, we really observe even in this 
case only a mean value; "point experiences” are never recorded. The 
assignment of symbols to experiences always contains then, strictly 
speaking, a statistical or, if we like, a collective element. Thus it is 
always a matter of making the assignment so as to go into detail to a 
greater or lesser degree. 

There can therefore never be the question— which the physicist 
who is infiuenced by the school philosophy often thinks must be 
asked— “Does strict causality hold in nature?” but rather, “What is the 
character of the correspondence between our experiences and the 
quantities describing the state of a system, which are subject to 
rigorous laws?” 

We see here, just as in the conception of the relativity theory, that 
the physicist, if he consciously or unconsciously maintains the stand- 
point of the school philosophy, is prevented from seeing the present 
physical theories as assertions about real physical experiences. He is 
easily led to find in them a mysterious, destructive element, raising 
philosophical difficulties. He may even find in them a contradiction 
to common sense. 

If we investigate more closely the nature of the epistemology of 
classical physics and its connection with school philosophy, we find the 
following: 

The general view was that, in the great system of symbols of which 
the physical theories are composed, there exists a framework which, 
with our experimental progress, must be gradually filled in. It appeared 


119 



modern science and its philosophy 

to be definitely established that all phenomena can be reduced to the 
motion of material points or the vibrations of a medium; that these 
material points at every instant possess definite positions and velocities 
by which all future states are unequivocally determined; that there are 
timpi variables with the help of which all phenomena can be most 
simply represented; and so on. It was believed that while it would be 
necessary to make many changes during the filling in of the framework, 
none would be made in the fundamental rods of the framework. 

Through the relativity theory and quantum mechanics this con- 
viction has been shaken. We know that even in those parts of the 
symbol system that form the framework, much has had to be changed 
and still more will have to be. We are in general no longer convinced, 
as we were formerly, that the parts of the symbol system forming the 
frame are already approaching a definitive form. This does not imply 
the acceptance of any skeptical standpoint, but only the rejection of * 
any distinction between the various portions of the symbol system. 

Every physicist is convinced that with experimental progress, with 
progressive refinement of the measuring technique, we will admit finer 
and finer structures and will always have to introduce new variables 
to describe the state of a system. Similarly, he must be convinced that 
there does not exist for all eternity a rigid framework which is char- 
acterized by the trio of space, time, and causality, and in which no ex- 
perience can change anything. He must be convinced rather that for 
these most general rules of assignment exactly the same must hold as 
for the more special rules, the dependence of which on the progress 
of human experience is not doubted. 

Classical physics led to the opinion that this framework was, in its 
essentials, completed. Hence it could be proclaimed by the school 
philosophy as eternal truth. Our modem theoretical physics, which 
admits progress in all parts of the symbol system, is skeptical only 
when viewed from the standpoint of the school phUosophy. From the 
standpoint of the purely scientific conception, which takes only ex- 
periences for granted and looks upon the symbol system constmcted 
for them as a means or an instrument, there is nothing skeptical in it. 
Would it show skepticism if we asserted that the ultimate machine for 


120 



twentieth-century physics and schooi philosophy 


traveling through space does not have to resemble the present airplane, 
not even in its most important parts; that it must have in common with 
the latter only one thing, the ability to fly? 

And now let us return to the question put at the beginning of this 
paper: What is the significance of the physical theories of the present 
day for the general theory of knowledge? From the standpoint of the 
school philosophy they signify a disintegration of rational thought; 
hence they are only rules for representing the results of experiments 
and not cognitions of reality, which are reserved for other methods. For 
those, however, who do not recognize these nonscientific methods, the 
present physical theories strengthen the conviction that even in ques- 
tions such as those concerning space, time, and causality, there is 
scientific progress, along with the progress in our observations; that it 
is therefore not necessary beside the thriving tree of science to assume 
a sterile region in which reside the eternally insoluble problems in the 
attempted solution of which men have been only rotating about dieir 
own axes for centuries. There are no boundaries between science and 
philosophy, if one only formulates the task of physics in accordance 
with the doctrines of Ernst Mach, using the words of Carnap: 

To order the perceptions systematically and from present perceptions to 
draw conclusions about perceptions to be expected. 


121 



CHAPTER 


5 


is there a trend today toward idealism in physics? 

I T is generally recognized that modem exact science, the creation 
of which in the age of Galileo and Newton led to the great ex- 
pansion of our technical civilization, is distinguished from an- 
cient and medieval science by the fact that the psychic, anthropo- 
morphic elements are being eliminated more and more from science. 
In place of the medieval doctrines of “the most perfect orbit,” “the 
position appropriate to a body,” “the difference between celestial and 
terrestrial bodies,” and the like, we now have the mathematically 
formulable laws of Newton’s principles, in which only observ able ^nd 
measurable quantities occur. tThere is no doubt that the/physics of 
Galileo and Newton has created a gulf between body and mind 
which did not exist in the anthropomorphic, animistic science of the 
Middle Age^This separation of the two became unpleasant to those 
who were interested in a science that would account for the behavior 
not only of inanimate bodies, but of all bodies in nature, including the 
human body. Thus arose the problem of explaining the mind on the 
basis of mechanistic physics, a problem which many have discussed, 
but without ever making any real advance, and which is actually only 
an apparent problem. Its insolubility in this form, which was really 
obvious to all, led many scholars to have a deep dislike for mech- 
anistic physics and to derive a malicious pleasure from every dif- 


122 



,is there a trend toward Idealism in physics? 

ficiilty that the latter encountered. R. Ruyer is quite right when he 
says: 

That, basically, many scholars are tortured by the burden of this 
original sin, mechanistic physics, is strikingly shown by their reaction when- 
ever the mechanistic or quantitative conception of physics appears to sufFer 
a setback. The most philosophical minds, far from being disturbed by this 
setback, hope every time to find in it an opportunity for introducing once 
more the subjective. This was the case with the discovery of potential energy, 
gravitation, and the degradation of energy, as well as chemical afiBnity.^ 

One need not be surprised, therefore, if the recent revolutions in 
the field of theoretical physics, the creation of the relativity and 
quantum theories, were received by the scholars whom Ruyer called 
“the most philosophical minds” with the same feelings as the preced- 
ing theoretical overtumings, such as the degradation of energy. Indeed, 
today one can hardly open a periodical or book dealing with the de- 
velopment of our general scientific ideas without meeting such ex- 
pressions as "the end of the age of Galileo,” “the failure of mechanistic 
physics,” “the end of the hostility of science toward the spirit,” “the 
reconciliation between religion and science.” There is even a book on 
modern physics, by Bernhard Bavink, entitled “Natural Science on the 
Path to Religion.” * 

Some are of the opinion that the new physical theories of the 
twentieth century have brought about a change in the general con- 
ception of the world as important as that caused by0he physics of 
Galileo, which replaced the animistic conception of the Middle Ages 
by the mechanistic one of modem times. In the same way, the new 
physics is supposed to form a bridge from the “mechanistic world con- 
ception” of the eighteenth and nineteenth centuries to the “mathe- 
matical conception” of the twentieth century. The latter is thought to 
be nearer, in a certain sense, to the medieval, animistic conception 
than to the mechanistic one, because in mathematics there resides an 
“ideal” or “spiritual” element, and a “mathematical world” is not as 

^R. Buyer, Revue de Synthdse 6, 187 (1933). 

* B. Bavink, Die Naturwissenschaft auf dem Wege zur Religion (Frankfurt am 
Main: M. Diesterweg, 1933). 


123 



modern science and its philosophy 


i 


foreign to the spirit as is a mechanical world. This view was pre- 
sented in all solemnity by General Smuts in his opening address at the 
celebration of the centenary of the British Association of Science on 
September 23, 1931.“ He said, among other things: 


There is the machine or mechanistic world view dominant since the 
time of Galileo and Mewton, and now, since the coming of Einstein, being 
replaced by the mathematician’s conception of the universe ... If matter 
is essentially immaterial structure or organization, it cannot fundamentally 
be so di£Eerent from organism or life ... or from mind, which is an active 
organize^ 


First, we must ask from the standpoint of the logic of science 
whether the physical theories of the twentieth century really contain 
any spiritualistic elements and, second, we must ask with what processes 
outside of physics the demand for a spiritualistic conception of nature 
is generally found to be associated. Let us begin by touching briefly on 
the second question in order to be able to consider the first in greater 
detail. 

It is certainly no accident that the culmination of the mechanistic 
conception of nature, as it is found, say, in the work of Laplace, 
coincided with the triumph of the French Revolution. It is certainly 
no accident that since that time the struggle against the “ideas of 
1789” has almost always coincided with a criticism of this conception 
of nature, a longing for a more idealistic or spiritualistic theory. The 
struggle against the “ideas of 1789” has been crowned in recent years 
by the fact that in a series of countries, especially in Italy and Ger- 
many, a directly opposite world conception has prevailed politically. 
This conception has a philosophic basis that is in sharp contradiction 
to the mechanistic conception of nature and urges a more “organismic” 
picture of the world, by which is meant a partial return to the spiritual- 
istic or animistic doctrines of the Middle Ages, just as the new con- 
ception of a state is connected with that of the Middle Ages. The ad- 
herents of this antimechanistic, organismic conception of nature strive 
to show that in the exact sciences “spontaneously,” “from purely 
scientific considerations,” a revolution has taken place. According to 

^Nature 128. 521 (1931). 

« 


} 



,is there a trend toward idealism in physics? 

^eir argument, on the basis of the relativity theory and quantum 
mechanics one can set up a conception of nature in which the mind 
again plays a role, and which is compatible with an “antimechanical, 
organismic, independent” biology^ 

As a typical example, a work of B. Bavink may be quoted. Bavink 
has a thorough knowledge of physics and biology and is an outstanding 
representative of the “organismic conception of nature.” He maintains 
that 

today there reigns within the circle of the sciences a willingness to tie once 
more the threads of science to all the higher values of human life, to Cod 
and the soul, freedom of will, etc.— threads that seemed almost completely 
severed; it is a willingness the like of which has not been present for cen- 
turies. That this change should take place at the present time is a coincidence 
almost hordering on the miraculous, for this change has in itself nothing to 
do with pohtical and social transformations; it manifestly arose from purely 
scientific motives.^ 

Whether or not the last sentence is correct is exactly the question we 
are trying to answer. 

On the other hand, in Russia, since the founding of the Soviet 
Union, a system has been established that seeks its philosophic basis 
in the "dialectical materialism” of Karl Marx as adapted by Lenin. I 
do not wish here to discuss the relation between this “dialectical ma- 
terialism” and what one is accustomed to call “materialism” in Germany 
and France. I want only to call attention to the fact that in countless 
articles in the philosophical and political journals of present-day Russia, 
the tendency toward spiritualism which is often found as an accompani- 
ment of modem physical theories is interpreted as one of the “phe- 
nomena of decadence” of science in capitalist countries.' In these 
articles the following line of thought oRen appears: in western 
Europe science, to be sure, is still making progress on individual prob- 
lems, such as the formulation of laws of atomic processes, just ias 

*B. Bavink, “The Sciences in the Third Reich" (in German), Unsere Welt 
25, 225 (1933). 

' As a recent example may be cited A. K. Timiriazew, “The Wave of Idealism 
in Modem Physics in the West and in Our Country” (in Russian), Pod znamenem 
marksizma, 1933, no. 5. 


125 



modern science and its philosophy ^ 

capitalist economy is still progressing technically. However, just as the 
life of the industrial population is shaken more and more by crises 
which finally make a generally acceptable solution impossible, so 
science, in spite of its progress in details, cannot produce a satisfying 
general picture of the processes in nature. In working on such a gen- 
eral picture it no longer proceeds scientifically, in the modem sense. 
Rather, it borrows from the animistic, spiritualistic physics of the 
Middle Ages and interprets modem theories in this light, because the 
intellectual trends dominant in political life obscure the scientific 
theories by wrapping them in a spiritualistic fog. 

In spite of the tremendous conflict between the “materialistic” 
Soviet Union and the states based on the organismic conception of the 
world, they all agree on the fact that the tendency toward spiritualism 
in modem physics corresponds to the ideology of the new organismic 
state. By some this tendency is welcomed as a necessary consequence 
of modem physics; by others it is condemned as an adulteration of 
it. That the representatives of both groups comprehend physics in this 
way is a fact that is as well established empirically as the best observa- 
tions of experimental physics, a fact which we must therefore take into 
account in any consideration of modem physical theories. 

I wish to say at once that the result of our investigation will be as 
follows: In the process of eliminating the “animistic” nothing has been 
changed in the slightest by the modern physical theories. This process 
continues irresistibly forward as before. He who would interpret 
physics by means of “psychic factors” had at the time of the physics 
of Galileo and Newton the same justification as today. The role of the 
“psychic” has remained exactly the same. Hence, if there does exist 
today a greater tendency toward spiritualistic interpretation, it is con- 
nected only with processes that have nothing at all to do with the 
progress of physics. 

Cjhe arguments that are supposed to show that psychic factors play 
a greater role in modem physics than in the physics of Newton are of 
various kinds. In one group it is claimed that the role of the “observing 
subject” in the relativity and quantum theories can no longer be 
eliminated from physical statements, as was still the case in “classical 


126 



is there a trend toward idealism in physics? 

physics^jjfrhis argument is often put as follows: Whereas in classical 
physics expressions like “length of a rod” or “time interval between two 
events” assert something about “objective” facts, in the Einstein 
relativity theory such expressions have a meaning only if the observer 
to whom they refer is specified. One can only say, for instance, “This 
body has a length of one meter with respect to this particular observer.” 
It appears, therefore, that every physical statement possesses a psycho- 
logical constituent. In popular literature on the relativity theory writers 
often even go so far as to compare the various lengths of a rod for 
various observers with the optical illusion that arises when one draws 
two straight lines of equal length but places different ornaments at 
their ends, producing the illusion of different lengths. 


< > > < 

^This conception,^ in its "scientific” as well as its “popular” form, is 
based on^a complete misunderstanding' of the relativity theory^Vhere- 
^ever in the theory of relativity reference is made to an observer, a 
physical measuring instrument can be substituted. It is asserted only 
that the results of the measurement will be different according as the 
motion of the measuring instrument is different. But in this there is 
nothing psychological,Jat any rate not any more than in classical 
physics. The role of the observer is in both cases exactly the same: he 
merely substantiates the fact that in a certain instrument a pointer 
coincides with a division mark on a scale. For this purpose the state of 
motion of the observer himself is quite immaterial. In the theory of 
relativity, as well as in classical physics, it is assumed that such a sub- 
stantiation is£ob]ective,” that there can never arise any difference of 
opinion in connection with it. Naturally, it remains “subjective” in 
the sense that some observer is necessary for it. Here “objective” means 
“the same for all subjects,” or “intersubjective.’^ 

Similar considerations have also been associated with the quantum 
theory. According to this theory, as Heisenberg showed, the position 
and the velocity of a given particle can never be exactly determined 
simultaneously. If one makes use of an experimental arrangement that 


127 



modern science and its philosophy 

' 

allows the position to be measured very accurately, the exacqfeeasure- 
ment of velocity by means of the same experimental arrangement is 
impossible. It is then held that whereas in classical physics a statement 
about the position and velocity of a particle was a statement about an 
objective fact, without any psychologic elements, in the quantum 
theory one cannot speak of the position and velocity of a particle, but 
only of what is given by a certain measurement. Thus every statement 
about particles involves the observer himself, and hence contains a 
psychologic element. To this argument we must reply as I have just 
indicated in the discussion of the relativity theory, pointing out that 
in quantum mechanics, too, what matters is never the observer, but 
only the instruments of observation. The role of the human being as 
observer is limited here again to establishing whether or not a pointer 
on a scale coincides with a division mark. This observation, however, 
is regarded here, just as in classical physics, as something “objective,” 
or better, as something “intersubjective.” What one can learn from the 
relativity and quantum theories in this connection is only what is also 
given by a consistent presentation of classical physics: every physical 
principle is, in the final analysis, a summary of statements concerning 
observations, or, if one wishes to speak in a particularly physical way, 
concerning pointer readings. 

We have thus seen that the new role of “observer” in physics 
cannot be used in favor of a tendency toward a more spiritualistic 
conception of physics. There exists, however, a whole series of other 
arguments which are customarily used to establish the approach or 
“return” of physics to the “organic, idealistic” conception of nature. 
Such arguments run somewhat like this; “Quantum mechanics con- 
tains a teleological element,” or “The indeterministic interpretation of 
the quantum theory makes room for free will.” Here we shall not con- 
sider these questions, but shall speak of a still more general argument 
for the “spiritualistic character” of modem physics. This is an argu- 
ment which in recent years has been repeated so often and by scholars 
of such prominence that there is danger that many, through becoming 
accustomed to such lines of thought, will accept them as justified— 
indeed, as obvious. These ideas have perhaps received the most 


128 



>is there a trend toward idealism in physics? 

extensive dissemination in 150,000 copies of a book by the outstanding 
physicist and astrophysicist, }. H. Jeans. Jeans depicts the present 
situation in physics as follows: 

Today there is a wide measure of agreement, which on the physical side 
of science approaches almost to unanimity, that the stream of knowledge is 
heading towards a nonmechanical reality; the universe begins to look more 
like a great thought than like a great machine. Mind no longer appears as an 
accidental intruder into the realm of matter.® 

Jeans bases his opinion that nature is to be regarded as something 
“spiritual” essentially on the assertion that modem physics has shown 
that one cannot give any mechanical representation of natural proc- 
esses, although one can give a mathematical representation. He says: 

The efforts of our nearer ancestors to interpret nature on engineering 
lines proved equally inadequate . . . On the other hand, our efforts to in- 
terpret nature in terms of the concepts of pure mathematics have, so far, 
proved brilliantly successful.’ 

In the laws of mathematics, in contrast to those of mechanics, of 
machinery, Jeans sees, however, a spiritual element. If nature behaves 
according to mathematical laws, it must be the work of a mind which 
can create mathematics, like the human mind, but is more compre- 
hensive. Jeans is so strongly convinced that the movement toward 
idealism is connected with the present state of theoretical physics that 
he keeps in view the possibility that, with a change of the theories of 
physics, there may again develop a movement away from idealism. 
Thus he says in a later book: 

So far the pendulum shows no signs of swinging back, and the law and 
order which we find in the universe are most easily described— and also, I 
think, most easily explained— in the language of idealism. Thus, subject to 
the reservations already mentioned, we may say that present-day science is 
favorable to idealism . . Yet who shall say what we may find awaiting us 
roimd the next comer? ® 

®J. H. Jeans, The Mysterious Universe (Cambridge: The University Press, 
1930), p. 158. 

Ubid., p. 143. 

*J. H. Jeans, The New Background of Science (Cambridge: The University 
Press, 1933), p. 296. 


129 



modern science and its philosophy 


Similar views are presented by Sir Arthur Stanley Eddington, in 
his book The Nature of the Physical World." While there is in this 
book a great deal that is beneficial in furthering the understanding of 
modem physics and in bringing its results to a wide circle of readers 
through a concrete and lucid presentation, yet it has numerous sections 
which Eddington himself regards as bold interpretation of present-day 
physics to which many will perhaps take exception, and which, in my 
opinion, form obstacles to the task of fitting physics into a self- 
consistent picture of the processes of the whole of nature. Eddington, 
like Jeans, believes that these views are matters of faith and that it is 
impossible to force one by proof to accept them. Thai is certainly tme. 
However, what can be shown clearly, in my opinion, is that these 
idealistic views have nothing at all to do with modem physics. If 
anyone wanted to accept them, he could have done so just as well 
in connection with the physics of Galileo and Newton, which is not 
less "mathematical” than twentieth-century physics. 

The arguments of both Jeans and Eddington depend on the con- 
trast between a physics that reduces everything to mechanics (that 
of Galileo and Newton) and one that bases everything on mathe- 
matical formulas (the physics of Einstein and the quantum theory). 
How can one formulate clearly the distinction between a “mechanical” 
and a “mathematical” basis for the processes of nature? Newtonian 
physics reduces all phenomena to the equations of motion for mass 
points between which there act central forces, that is, to a system of 
differential equations. The mechanics of Einstein changes these dif- 
ferential equations in a few respects which give essential differences 
only for very high velocities, and points out that the equations so 
changed have mathematically the same form as the geodesics in a 
curved (non-Euclidean, Riemannian) space. In place of one system of 
differential equations, another occurs. Why then is one theory called 
“mathematical,” the other “mechanistic”? Surely similarity to the 
geodesics cannot be the only reason, for Newtonian physics can also 
be brought into this form without any difficulty. 

Adhering to the concrete interpretation of physics as a representa- 

• Cambridge: The University Press, 1928. 


130 



is there a trend toward Idealism in physics? 


tion of observable facts, we can try to summarize the difference be- 
tween “mechanical" and “nonmechanical, mathematical” physics ap- 
proximately as follows: By means of Newtonian mechanics we can 
describe the motions of bodies with which we deal in everyday life, 
so long as they have also the velocities that are encountered in daily 
experience. To this class of bodies belong the ordinary tools such as 
hammers and tongs, but also such things as steam engines, automobiles, 
and airplanes. During the reign of the physics of Galileo and Newton 
it was believed that in time it would be possible by means of these 
same equations to describe also the motions of the smallest particles of 
matter, such as atoms and ions, as well as the motions of celestial 
bodies during arbitrarily long time intervals and with arbitrarily high 
velocities. In other words, it was believed that all processes of nature, 
in the large and in the small, could be covered by the same laws that 
had been established for the motions of “bodies of average size with 
moderate velocities.” This belief has been shaken by the development 
of physics in the twentieth century. We know today that the motions 
of bodies with velocities comparable to that of light can be described 
only with the help of the relativity theory of Einstein, the motions of 
the smallest particles in the atoms only with the help of quantum and 
wave mechanics. 

If we understand by mechanics the doctrine of the motion of 
“bodies of average size with moderate velocities,” then we can rightly 
say that modem physics has established the impossibility of a mechani- 
cal basis for the processes of nature. If we say, however, that the 
mechanical foundation has been replaced by a mathematical one, it is, 
in my opinion, a very inappropriate mode of expression. We ought to 
say, rather, that the place of a special mathematical theory, that of 
Newton, has been taken by more general theories, the relativity and 
quantum theories. The opinion that a special mathematical theory 
could represent all the processes of nature has turned out to be false; 
that is all. But from this fact no contrast between the propositions 
“Newtonian physics = mechanics = materialism,” on the one hand, and 
“modem physics = mathematics = idealism,” on the other hand, can be 
deduced. . 


131 



modern science and its philosophy 


Newton, in his Mathematical Principles of Natural Philosophy,^" 
replaced the matter filling the world and acting tlirough pressure, col- 
lisions, and fiuid vortices, as pictured by the Cartesians, by small 
masses, almost lost in vast empty space and acting on each other only 
through forces at a distance. When this work was published, the new 
theory was hailed by many of his followers as a triumph over the 
materialism of the "Epicureans.” 

As proof, one need only read the famous controversy between 
Leibniz and Clarke,*^ in which Clarke defends Newton's teachings 
against the attacks of Leibniz. Clarke says in his first reply: 

Next to the corruptible dispositions of human beings, it [the disavowal 
of religion] is to be ascribed first of all to the false philosophy of the ma- 
terialists, who oppose the mathematical principles of philosophy [i.e., 
Newton’s] . . . These principles, and indeed only they, show matter and 
the body as the smallest and most insignificant part of the universe. 

Hence at that time Newton’s followers, in so far as they were 
adherents of spiritualistic metaphysics, extolled his teachings as “mathe- 
matical” and “spiritual” in contrast to materialism. Today those with 
analogous philosophic inclinations say that Newtonian physics was 
“materialistic,” but that Einstein has again brought in a “mathematical,” 
“spiritual” element in place of the mechanical one. 

We have already seen that the assertion that the laws of nature 
are not “mechanical” but "mathematical” means only that the laws are 
expressed not by means of the special mathematical formulas of 
Newton, but by means of the more general formulas of the relativity 
and quantum theories. When, however, we say, not that the formulas 
used to describe nature are mathematical, but that the world is mathe- 
matical, it is difficult to say what we mean. By mathematics, con- 
sidered concretely, we can only understand a system of formulas or 
propositions. With these formulas are to be correlated the observations 

“I. Newton, Philosophiae Naturalis Principia Mathematica (1687), tr. by 
A. Motte (1729), rev. by F. Cajori (University of California Press, 1934). 

A Collection of Papers, which Passed between the Late Learned Mr. Leibnitz, 
and Dr. [Samuel] Clarke, in the Years 1715 and 1716 . . . (London: J. Knapton, 
1717). 


132 



5 there a trend toward idealism in physics? 

that we make of the processes of nature, if the formulas are to repre- 
sent physical theories. The processes themselves, however, do not 
consist of these formulas. In an assertion such as “The world is basically 
mathematics,” the word “is” can only be used in a mystic sense, as it 
occurs perhaps in the sentence “This architecture or this music is pure 
mathematics.” 

In order to make his views clear, Jeans must speak of the world 
architect; he represents him, not according to Newtonian physics as a 
kind of engineer, but according to modem physics as a kind of mathe- 
matician. Since engineers also produce their work according to mathe- 
matical formulas, Jeans has to indicate the distinction between the 
engineer and the world creator somewhat as follows: The engineer fits 
his formulas to the observations, whereas the creator invents formulas 
at will and then constructs the world according to them. Jeans brings 
in here the difference between “pure” and “applied” madiematics. The 
engineer is an applied mathematician, the world creator a pure one. 
Jeans tries to show it in this way: The man who works in pure mathe- 
matics invents formulas and propositions without any regard to the 
question of practical application; later, it turns out that the physicist 
or engineer, by means of the results obtained by the pure mathe- 
matician, can represent the processes of nature, of which the pure 
mathematician knew nothing when he devised his theory. This can only 
be explained by saying that the processes are themselves the work of 
a pure mathematician, and the theoretical physicist who finds these 
formulas for representing observations is only rediscovering the ideas 
of the pure mathematician who created the world. The creations of 
the demiurge must accordingly agree to a large extent with those of a 
human pure mathematician. 

The assertion that the world is built according to the principles of 
“pure” mathematics is to be found not only in the works of Jeans. It 
is very often used in setting up mystical conceptions of the world. If 
one wants to be clear as to its meaning, one must first of all obtain 
clarity as to the meaning of the propositions of “pure” mathematics in 
general. According to the conception of B. Russell and L. Wittgenstein, 
which is also that of the Vienna Circle, the propositions of pure mathe- 



modern science and its phiiosophy 


matics are not statements concerning natural processes, but are purely 
logical statements concerning the question of what assertions are equiv- 
alent to one another, which can be transformed into one another by 
formal transformations. The propositions of pure mathematics, there- 
fore, remain correct, no matter what the natural processes may be; these 
propositions can be neither confirmed nor refuted by observation, since 
they state nothing concerning the real processes of nature. Mathemati- 
cal theorems, as is often said, are of an analytic character. 

For example, if I prove the theorem “The sum of the angles of a 
triangle is equal to 180“” as a proposition of pure mathematics, I 
prove only that from the axioms of Euclidean geometry, including the 
axiom of parallels, it follows by logical transformation that the sum of 
the angles of a triangle is equal to 180“ if the straight lines and points 
of which it consists have all the properties ascribed to them by the 
Euclidean axioms. That is to say, if for a concrete physical triangle 1 
can establish by observation the validity of the Euclidean axioms, then 
the sum of the angles is equal to 180“. In other words, the statements 
"The sum of the angles is 180“” and "The axioms are valid” are only 
two expressions of the same thing, two statements with the same con- 
tent (where, of course, the proposition of the sum of the angles is 
only a part of the content of the whole system of axioms). Once this 
has become clear, the world, whatever it may be, will always obey the 
propositions of pure mathematics; the assertion that it obeys them says 
nothing at all about the real world. It says only what is self-evident, 
that all statements about the world can be replaced by equivalent 
statements. 

Something else must obviously be meant, however, when Jeans and 
so many others say that the world is constructed according to the 
principles of “pure” mathematics. As an example, the following is ad- 
duced: Mathematicians— ChristofFel, Helmholtz, Ricci, Levi CivitA, and 
others— long ago built up the theory of the curvature properties of 
Riemannian space. When Einstein set up his general theory of rela- 
tivity, he found this whole branch of mathematics ready for him. Al- 
though it was invented without any intention of its being used in 


134 



is there a trend toward idealism in physics? 


physics, Einstein was able to apply it in his theory of gravitation and 
general relativity. One must therefore assume that the creator built 
the world according to those principles of pure mathematics. Other- 
wise it would be an inconceivable coincidence that such a complicated 
branch of mathematics, developed for quite other purposes, could be 
used for the theory of gravitation. 

We have already seen that this assertion cannot mean that the 
^rld is built according to the propositions of the Riemannian cuirva- 
ture theory or of the absolute difiFerential calculus invented by Ricci 
and Levi Civiti; for these propositions, like the proposition of the 
sum of the angles and all other propositions of pure mathematics, are 
only statements of how one can express the same thing in different 
ways. The assertion, therefore, can only mean that the concepts and 
definitions of pure mathematics— the geometry of Riemannian spaces 
—created certain structures— the Christoffel three-index symbols, the 
Riemannian curvature tensor— which could be utilized in the Einstein 
theory of gravitation. This, however, is the same, although perhaps 
on a higher level, as saying: ‘The concepts of the square or the square 
root or the logarithm have come out of pure mathematics; it is therefore 
amazing that they also occur in the formulas of physics.” If we now use 
the possibility of representing the world according to Einstein, with 
the help of the Riemannian curvature tensors, as proof that the world 
was created by a mathematician, we might have said with the same 
justification, back in the time of Newton, that the world must have been 
created by a mathematician; for in Newton’s formulas the cliief role 
is played by the “square of the distance,” and the concept of the 
square of a number originated in geometry and was introduced with- 
out any regard for physics. If we consider the matter from this stand- 
point, that is, if we speak not of mathematical theorems but of mathe- 
matical concepts, a little reflection shows that the distinction drawn 
by Jeans between engineer and mathematician, or between “applied” 
and “pure” mathematics, cannot be maintained. 

As a matter of fact, concepts such as those of Riemannian curvature 
have always been invented for the purpose of representing some prob- 


135 



modern science and its philosophy 


I 


lem of concrete reality, for describing processes of nature. The con- 
cepts of Riemannian geometry all go back to the problem of describing 
the motion of a real rigid body in general coordinates; one need only 
recall the work of Helmholtz on the facts at the basis of geometry 
Riemann, Christoffel, and Helmholtz set up certain mathematical ex- 
pressions which are equal to zero in the case of the motion of a rigid 
body, according to the usual laws of physics. When Einstein proceeded 
to formulate the deviations from these laws, it was clear that he had 
to begin with the expressions that gave the properties of rigid bodies, 
according to classical physics, in a form valid for all coordinate sys- 
tems. If there existed any deviations expressible independently of the 
coordinate system, it had to be possible to express them so that the 
quantities which in the old physics had the value zero were now dif- 
ferent from zero and took on values depending in a simple manner 
on the distribution of matter. If such a simple dependence did not exist, 
then there could exist no laws independent of the coordinate system, as 
Einstein required. If such laws did exist, it had to be possible to ex- 
press them through the concepts that were at hand for representing 
the motion of a rigid body. But this gives no evidence that the world ' 
creator was a “pure mathematician.” The' only thing that must be re- 
garded as a real and astonishing characteristic of nature is the fact that 
there do exist, in general, simple laws for the description of nature. 
This, however, has nothing to do with the distinction between “me- 
chanical” and “mathematical.” 

If today expressions with spiritualistic coloring are used to a greater 
extent than in the nineteenth century, this has no connection with any 
“crises in physics” or with any “new physical conception of nature.” It 
is rather associated with a crisis in human society arising from quite 
different processes. In opposition to the materialistic social theories 
there have come into the foreground movements based on an “idealis- 
tic” picture of the world. These movements seek support in an idealistic 

“ “XJber den Urspning und die Bedeutung der geometrische Axiome” ( 1870), 
translated as “On the Origin and Significance of Geometrical Axioms," in Helm- 
holtz, Popular Lectures on Scientific Subjects, series 2, tr. by E. Atkinson (London: 
Longmans, Green, 1881). 


136 



^ is there a trend toward idealism in physics? 

or spiritualistic conception of nature. Just as at the end of the nine- 
teenth century analogous movements made use of energetics, the 
electromagnetic picture of matter, and so on, to prophesy the end of 
“materialistic” physics, so today die relativity and quantum theories 
are being used. All this, however, has no real connection with the prog- 
ress of physics. 


137 



CHAPTER 



mechanical "explanation” or mathematical description? 

% A #HEN the physics of Galileo and Newton put an end to 
\m\mche animistic period, it received the honorary title “mathe- 
y Viatical.” At that time Newton wrote the Mathematical Prin- 
ciples of Natural Philosophy. A pronounced antimaterialistic tend- 
ency was ascribed to this work, because in it palpable impacts of 
masses upon one another were replaced by a pure mathematical for- 
mula, the law of attraction. The physics of Galileo and Newton did 
not come to be called “mechanistic” until it had become customary 
to use the treatment of mechanics as a model for every other field. 

The Newtonian mechanics was looked upon more and more as the 
standard type of a theory. Helmholtz, for example, said: 

To understand a phenomenon means nothing else than to reduce it to 
the Newtonian laws. Then the necessity of explanation has been satisfied in a 
palpable way. 

It was completely forgotten that in Newton’s own day his theory 
was looked upon as a set of abstract mathematical formulas, which 
needed a mechanical explanation to satisfy man’s desire for causality. 
Newton himself recognized this need, but he declined to take part in 
satisfying it himself, when he made the now well-known remark, 
"Hypotheses non fingo.” But men like Huyghens and Leibniz never 
considered the Newtonian theory a physical explanation; they looked 


138 



meohanical "explanation" or mathematical description? 


upon it as only a mathematical formula. Thus even at the very begin- 
ning of the development of mechanistic physics, it was not easy to de- 
fine precisely the distinctions between a “mechanistic” and a “mathe- 
matical” theory. 

Much later a contradiction began to be noticed between “mathe- 
matical” and “pictural,” ^ and it was asserted that only the reduction 
to mechanics guaranteed a theory that provides a pictural representa- 
tion and that without this pictural character no real understanding was 
possible. It was claimed, for example, that Maxwell’s field equations of 
electrodynamics are not pictural, if they are not illustrated with a 
mechanical model. 

There were two motives involved in the resistance to the abandon- 
ment of mechanism. First, there was no inclination to renounce the 
“explanatory” value that was ascribed to mechanistic theories only; 
second, it was feared that the abandonment of mechanistic explana- 
tions would lead to a return to medieval animistic anthropomorphic 
science. 

But in speaking of this desire for picturization we ought to be , 
clear as to the actual meaning of the term. The mechanical laws de- 
scribe for us the ordinary experiences of daily life— the use of tools, 
automobiles, firearms, as well as the movements of the planets. We find 
it desirable to interpret all other experiences by analogy with those 
that are most familiar. 

The physics of the nineteenth century showed that this wish can- 
not be fulfilled. Electromagnetic phenomena cannot be reduced to the 
same mechanical laws that govern guns or tools. Nevertheless, the laws 
concerning electromagnetic phenomena may be considered in a 
broader sense as also pictural. We can verify experimentally the valid- 
ity of such physical laws as, for example. Maxwell’s field equations if 
we can derive from them a result that is directly observable experi- 
mentally. The experiment consists in observing the position of the 
pointer of an ammeter or of some other measuring apparatus, or the dis- 

^We here translate the German word amchaulich by the English pictural, 
which we take to imply, as the German does, both the ordinary visual perception 
of objects and the mental visualization oC them. See also Chapter 7. 

139 



modern science and its philosophy 


placement of some colored spot. But these are precisely the kind of 
observation that we use to verify the laws of ordinary mechanics. In 
the end only gross mechanical events, which certainly are pictural, 
are derived from the equations of the electromagnetic field, and this 
holds in all physics, including that of the twentieth century. In this 
sense physical laws must be pictural if they are to have any scientific 
meaning at all, for otherwise they are not experimentally verifiable. 
The fundamental equations by themselves need not be pictural, since 
they cannot be submitted to any direct experimental verification. 

For a long time efforts were made to set up a mechanical explana- 
tion of Maxwell’s theory. There was always the feeling that without it, 
something essential to the understanding of electromagnetic phenom- 
ena was missing. Heinrich Hertz finally cut the Gordian knot, so to 
speak, when he said: “Maxwell’s theory is nothing else than Max- 
well’s equations. That is, the question is not whether these equations 
are pictural, that is, can be interpreted mechanistically, but only 
whether pictural conclusions can be derived from them which can be 
tested by means of gross mechanical experiments.” 

These words gave birth to what we call today the “positivistic con- 
ception” of physics. Positivistic physics thus replaced mechanistic 
physics. The mechanistic explanation could now be abandoned as a 
foimdation, without at the same time renouncing the achievements of 
the epoch of Galileo and Newton. If a positivistic conception of physics 
was accepted, the rejection of medieval animism was as complete as in 
mechanistic physics. In place of the mechanical model there was the 
mathematical formula with its experimentally verifiable results. In this 
sense, it may be said that the positivistic conception replaced the 
mechanistic interpretation with a mathematical one. Before anything 
was known about the relativity or the quantum theories, before, there- 
fore, even the “rebirth of idealistic physics in the twentieth century,” 
Hertz, Mach, Duhem and others had already seen that the essential 
point in every explanation of nature is not the mechanical model but 
rather the construction of mathematical relations. 

The historical error is often made of connecting the struggle of Mach 
and Duhem for the positivistic physics with their aversion for atomism. 


140 



mechanical "explanation" or mathematical description? 

so that a victory for atomism was considered a defeat for positivism. 
In reality the champions of atomism, Maxwell and Boltzmann, were 
exactly of the same opinion concerning the general nature of a physical 
theory as Hertz and Mach. The difference in their views about the 
value of atomistic theories arose only because they differed in their 
estimates of the convenience with which the actually known physical 
phenomena could be derived from these theories. 

A few quotations from the writings of Maxwell and Boltzmann 
win at once make clear what they thought concerning the structure of 
physical theories and their connection with experience. 

In the introduction to his treatise. On Faradays Lines of Force, 
Maxwell expressed himself quite clearly on these questions. Boltzmann 
says in the notes to his German translation of this paper: 

Maxwell’s introduction proves that he was as much of a pioneer in the 
theory of knowledge as he was in theoretical physics. All the new ground in 
the theory of knowledge that was broken in the next forty years, is already 
clearly marked out in these few pages; indeed the very ideas are illustrated. 
Later epistemologists treated all this more fully, but also with greater bias. 
They set up rules for the future development of a theory after it had 
already developed and not before, as was the case with Maxwell. 

Maxwell describes how he first found a convenient mathematical 
formulation of the laws of electricity and magnetism that were already 
known, and then used the mathematical concepts so created for the 
construction of the new laws; 

In order therefore to appreciate the requirements of the science, the 
student must make himself familiar with a considerable body of most intricate 
mathematics, the mere retention of which in the memory materially inter- 
feres with further progress. The first process therefore in the effectual study 
of the science, must be one of simplification and reduction of the results of 
previous investigation to a form in which the mind can grasp them. The 
results of this simplification may take the form of a purely mathematical 
formula or of a physical hypothesis.^ 

® J. C. Maxwell, “On Faraday’s Lines of Force,” Transactions of the Cambridge 
Philosophical Society 10, part 1 (1855); Scientific Papers of James Clerk Maxwell 
(Cambridge University I^s, 1890), vol. 1, p. 155. 


141 



modern science and its philosophy 


■ But Maxwell by no means tliinks that there is an essential antithesis 
between a “purely mathematical formula” and a “physical hypothesis.” 
He judges them only according to their practical exchange value, and 
finds that each of them has its advantages and disadvantages. Hence 
he looks for a kind of theory that is more general than either and that 
combines the advantages of both without the disadvantages. This more 
general sort of theory Maxwell finds in the physical analogy. It com- 
prehends both the physical hypothesis, in which an analogy is drawn 
between electromagnetic events and a mechanical model, and a mathe- 
matical formula, which points out an analogy between the phenomena 
of electricity and certain mathematical relationships that are given 
by the electromagnetic-field equations. 

Maxwell continues: 

In the first case [of a purely mathematical formula] we entirely lose sight 
of the phenomena to be explained; and though we may trace out the conse- 
quences of given laws, we can never obtain more extended views of the con- 
nections of the subject. If, on the other hand, we adopt a physical hypothesis, 
we see the phenomena only through a medium, and are liable to that blind- 
ness to facts and rashness in assumption which a partial explanation en- 
courages ... In order to obtain physical ideas without adopting a physical 
theory we must make ourselves familiar with the existence of physical 
analogies. By a physical analogy I mean that partial similarity between the 
laws of one science and those of another which makes each of them illus- 
trate the other. Thus all the mathematical sciences are founded on relations 
between physical laws and laws of numbers, so that the aim of exact science 
is to reduce the problems of nature to the determination of quantities by 
operations with numbers. Passing from the most universal of all analogies to 
a very partial one, we find the same resemblance in the mathematical form 
between two different phenomena giving rise to a physical theory of light.® 

In the introduction to his lectures on the theory of gases Boltzmann’s 
view is brought out very clearly. In one place he says: 

The question as to the fitness of the atomistic philosophy is naturally 
wholly untouched by the fact stressed by Kirchhoff that our dieories about 
nature bear the same relation to it as symbols to the things symbolized, as 
letters to sounds or as musical notes to tones, and by the question whether 

®Iiid.,pp. 155 f. 


142 



me^anical "explanation" or mathematical description? 


it may not be expedient to consider our theories as pure descriptions so 
that we may always recall their relation to nature. Therefore the question is 
whether the pure differential equations or atomism will one day turn out 
the more complete descriptions of phenomena.^ 

^L. Boltzmann, Vorlesungen iiber Castheorie (Leipzig; Barth, ed. 3, 1923), 
vol. 1, p. 6. 


143 



CHAPTER 



modern physics and common sense 


A LMOST every new physical theory has to face the commonplace 
accusation that it stands in contradiction to everyday experi- 
# \ ence or, as it is sometimes put, that it contradicts common sense. 

The heliocentric system of Copernicus is perhaps the most famous 
example of a theory that has been charged with being in contra- 
diction to the evidence of our senses. For on the one hand, many of 
us have grown accustomed to believing that our eyes disclose to us the 
earth is at rest with the sun and planets revolving around it. On the 
other hand, the Copemican theory, established since the time of Galileo, 
maintains that the sun is at rest and the earth revolves around it, "in 
contradiction to this immediate testimony of our eyes.” 

Although this accusation has been made in innumerable books, ar- 
ticles, and lectures, it is difficult to understand what could be meant 
by a physical theory’s being in contradiction to the evidence of our 
senses. Such a contradiction could occur only if the theory implied 
propositions about directly observable matters which were not con- 
firmed by actual observations. This, however, is not the case for the 
Copemican theory. In point of fact, the content of our observations 
on sunsets and sunrises can be deduced from the heliocentric theory 
as well as from the geocentric one. It must therefore be noted that 
our sensory observations are not adequately formulated in such a 
proposition as “We see immediately that it is the sun which is moving 


144 



modern physics and common sense 


and not the earth.” For when we formulate our direct observations of 
die solar motions we obtain statements like the following: “The 
distance between the sun and the horizon increases from morning to 
noon, and decreases from noon to evening.” And such propositions 
can obviously be deduced from the Copemican as well as from the 
Ptolemaic theory. Consequendy, the charge of a contradiction between 
a given physical theory and the evidence of our senses cannot be sup- 
ported, if we understand by “evidence of our senses” the propositions 
that can be deduced from the theory and checked by experiment. 

Our preference for the heliocentric theory rather than the geo- 
centric one is based on the following two points. First, the heliocentric 
theory makes it possible to calculate more simply the observable posi- 
tions of the planets in their orbits than does the geocentric theory. 
Second, generalizations, such as Newton’s laws of motion, can be made 
much more easily on the basis of the heliocentric theory than on the 
geocentric one. Such facts of observation as the perturbations of the 
planetary orbits could be shown to be consistent with the geocentric 
theory only by using formulas and computations that would exceed 
the powers of a human mathematician. On the other hand, starting 
from Newton’s laws of motion these perturbations can be calculated 
with the help of relatively simple formulas within die framework of 
the heliocentric theory. 

So much for the logical issues involved in the charge. There is, 
however, an obvious psychologic reason why people continue to 
make it. If we consider only the facts of everyday experience, such as 
the daily rising and setting of the sun, moon, and stars, we discover 
that they can be deduced from the geocentric theory with great ease. 
To deduce from this theory the facts concerned with the movements of 
the planets is still possible, though highly inconvenient and involved. 
If now we also consider the phenomena of perturbation, we find that 
to deduce them from the geocentric theory is practically impossible, 
and that the Copemican theory alone is capable of handling them. On 
the other hand, the deduction of everyday phenomena like sunrise 
and sunset can be performed more easily and simply from the geo- 
centric theory than from the more “artificial” heliocentric theory— es- 


145 



modern science and its philosophy 


pecially if we take the latter in its completely developed form as given 
by Newton, so that it becomes adequate for encompassing the phe- 
nomena of planetary perturbation. 

We may therefore formulate the difference in function of the two 
theories in the logical system of science as follows: If we are concerned 
J with the deduction of phenomena occurring in the narrow domain of 
everyday experience, it is convenient to use the geocentric theory. But 
since the scope of the facts falling within the province of this theory is 
very restricted, if we seek a theory that will embrace as many phe- 
nomena as possible (including phenomena remote from daily experi- 
ence), we must make use of the heliocentric theory. 

Because the geocentric theory is more convenient for handling 
familiar phenomena of daily life, it is often called an “intuitive” or 
“visualizable” theory. Indeed, many people have the impression that 
they fully understand this theory by a direct, immediate observation 
of facts, and that no logical gap separates it from the content of our 
naive sensory observations. However, all these characterizations of 
the theory mean nothing more than that a relatively simple chain of 
deductions connects it with the observable phenomena of daily life. 

On the other hand, because the deduction of our most familiar 
observations from the heliocentric theory is not as simple as in the 
case of the geocentric one, there is a common impression that the 
former is not connected closely with everyday experience. It has 
therefore been called an “abstract” or “nonintuitive” theory, appar- 
ently because the facts of experience upon which it rests occur only in 
the context of scientific research, as when we use astronomical in- 
struments to determine the exact positions of the stars. 

This familiar distinction between “intuitive” and “abstract” theories 
is, however, often misunderstood and misused. Thus, it has been said 
that “intuitive” theories are the only sound ones, because they alone 
serve as a stimulus to the physicists to carry on fruitful research. They 
have also been said to offer a true representation of the real world in 
terms of faithful pictures or images, thus furthering the scientist in his 
search for the truth. On the other hand, "abstract” theories, which 


146 



, modern physics and common sense 

employ mathematical formulas rather than the method of pictorial 
representation, have been characterized as impediments to scientific 
progress. It has been said of them that they can at best register knowl- 
edge already achieved, but that they are of no help in the discovery of 
new knowledge. 

These alleged differences between intuitive and abstract theories 
are regarded as particularly significant by followers of Kant and other 
adherents of German idealistic philosophy. For intuitive or visualiz- 
able theories have been claimed to be “anschaulich” in the sense 
specified by Kant. Now in everyday German the word Anschauung 
simply means the observation of an object, especially a visual obser- 
vation, for example, the visual perception of a table. But for Kant and 
other idealistic philosophers it frequently has a half-mystical meaning. 
Thus, when they speak of an “innere Anschauung” or “internal intui- 
tion” of the properties of a triangle, for example, they do not mean to 
refer to a visual perception of a triangle, but to some mental act of 
concentration upon the triangle, which itself exists only in the mind 
as an image. We shall translate the term innere Anschauung by the Eng- 
lish word “picturization,” and anschaulich by “pictural.” According to 
the Kantian philosophy, we are supposed to be able to “see” directly 
in this mental act such things as that the sum of the angles of a triangle 
must be equal to two right angles. Accordingly, this alleged power of 
picturization is claimed to be able to establish the essential properties of 
triangles without appealing to empirical tests and without regarding the 
theorems of geometry as propositions consisting of observation terms 
—where “observation” is used in the familiar sense of observation 
through the senses. In this way Euclidean geometry was “proved” to Jbe 
the only legitimate geometry. The non-Euclidean geometries were al- 
leged to have been demonstrated to be false, because their theorems 
were not guaranteed by picturization. In opposition to these dicta, we 
must note, however, that the propositions of non-Euclidean geometry 
can be formulated in observational terms just as the Euclidean prop- 
ositions can be, and that the alleged absurdity of the former arises 
only from the fact that they are required to be “tested” by an internal 
intuition. 

-» • 


147 



modern science and its philosophy 


These ambiguities in the meaning of “intuition” or Anschauung 
have made it possible for the intuitive theories of physics to acquire a 
special importance in the eyes of philosophers and even of some physi- 
cists. In the Germany of 1939, for example, the differences between 
intuitive and abstract theories were linked up with differences between 
races. Intuitive theories were associated with the characteristics of the 
Nordic race, while the abstract ones were taken as expressing the pe- 
culiarities of the Mediterranean races, especially of the Semites, though 
sometimes even of the French. Naturally, since the Nordic race was re- 
garded as the superior group, intuitive theories were also superior to 
the abstract ones. 

But the preceding discussion of the differences between the helio- 
centric and the geocentric theories makes clear just what the difference 
between these two types of theory really signifies. What is usually 
called an "intuitive” or pictural theory is simply a theory very well 
adapted to formulate in a simple and convenient manner our everyday 
j experiences; an abstract theory, on the other hand, attempts to cover a 
more inclusive domain of phenomena, and does not hesitate therefore 
to formulate the familiar facts of daily life in a somewhat complex 
marmer. 

If we are once clear about the differences between the geocentric 
and heliocentric theories, we are prepared to understand and evaluate 
the objections often raised against more recent physical theories, such 
as the theory of relativity or the quantum theory. These objections tend, 
in the main, to allege that the new theories are too abstract and that 
they use conceptions very far removed from those employed in every- 
day experience. Indeed, some physicists have argued that the newer 
conceptions are too remote from the notions employed by physicists 
themselves in their laboratory researches. 

Our previous analysis supplies us with a clue for evaluating these 
objections. The laws formulated by so-called classical physics are 
obtained from the study of material bodies and optical phenomena as 
these occur either in everyday experience, or in physical experiments 
carried on under conditions not very different from die circumstances 


148 



modern physics and common sense 


of everyday experience. With the help of these laws, the behavior of 
bodies and of light in these domains can be expressed in a simple and 
useful way. Until the end of the nineteenth century these laws have 
been supposed to govern all the phenomena of the physical world, 
from the smallest bodies, like electrons, to the largest ones, like suns 
and stars, and from motions with low velocities to motions with the 
velocity of light. But the development of physical research at the end 
of the nineteenth century and the beginning of the twentieth showed 
that the laws of classical physics are valid only for small velocities and 
large bodies— to speak more exactly, for masses far greater than the 
mass of an electron, and for velocities far less than the velocity of 
light. 

Now to express the behavior of things that do not occur in daily 
experience, such as small particles with large velocities, we require a 
generalization of classical physics and new types of physical law. 
These new laws are naturally more complex than the laws of classical 
physics, especially if we compare them in their applications to phe- 
nomena of common experience; in these domains the laws of classical 
physics are limiting cases of the new laws of relativity and quantum 
physics. On the other hand, there is no sense in saying that these new 
laws, when applied to phenomena, such as those within the hydrogen 
atom that do not occur in daily experience, are more complex than the 
classical laws, because the classical laws do not apply to these phe- 
nomena at all. Consequently, only in the sense indicated is it correct 
to say that the laws of classical physics are more closely related to 
everyday experience than are the laws of twentieth-century physics, 
namely, the laws of relativity and quantum physics. 

It is evident, therefore, that from the standpoint of the logical 
analysis of science the difference between intuitive and abstract theo- 
ries is a very superficial one. That distinction does not play the im- 
portant role in science that philosophers and laymen influenced by 
orthodox philosophy suggest it does. 

A glance at the history of science also reveals that the alleged wide 
gap between these two kinds of theory is not understood in the same 


149 



modern science and its philosophy 


manner at every period of the history of science. Thus, at the time of 
its discovery, Newton’s theory of motion was regarded as an abstract, 
merely mathematical theory; in our own day, however, it is often cited 
as an example of an intuitive theory, especially when philosophers wish 
to establish the abstract character of the theory of relativity and of 
quantum mechanics by contrasting them with an intuitive theory. In 
truth, however, the alleged diflFerence between Newtonians and rela- 
tivistic mechanics depends only on the undeniable fact that the difficult 
and complicated calculations and deductions required for under- 
standing Einstein’s theory of relativity are not required for under- 
standing the phenomena of everyday experience. For these phenomena 
can be formulated with the help of the more simple Newtonian theory 
of mechanics. Hence the statement that a theory like Einstein’s is 
abstract and nonintuitive simply means that it is more complex than 
is necessary for a theory that need describe only the facts of daily 
experience. 

It is often maintained that whether a given scientist will be more 
inclined to use an intuitive rather than an abstract theory is a function 
of his personality. These psychologic factors are sometimes taken to 
be facts of individual psychology, sometimes of race or of nationality. 
It seems to me, however, that the importance of such psychologic 
considerations has been exaggerated. Such psychologic factors seem 
to play little role in the work of the great masters of science to whom 
we are chiefly indebted for the present state of the sciences. Some 
masters of science exhibit a “double personality.” 

For example, the great English physicist J. Clerk Maxwell em- 
ployed elaborate mechanical models in his work, as well as abstract 
mathematical theories. He discovered many new facts with the help of 
both types of theory, and was himself firmly convinced that there is no 
fundamental difference between these two types from the point of view 
of the logical analysis of science. Indeed, he emphasized the fact that 
both abstract and intuitive theories are special cases of a more inclu- 
sive type of formulation of physical facts, namely, formulation or 
representation by analogies. And if we examine the work of Boltz- 
maim and other great physicists interested in the logical structure of 


150 



, modern physics and common sense 

science, we find them regarding physical theories in the same charac- 
teristic way. 

"We may therefore sum up our considerations on the differences 
between abstract theories and those alleged to be intuitive and com- 
prehensible to common sense. Intuitive theories try to preserve so far 
as possible the method of representation invented for handling 
most simply the facts of daily life. Such theories are therefore com- 
prehensible to everyone who is concerned with those facts and are thus 
comprehensible to so-called common sense. Mechanical images or 
models are particularly cherished in this context, and conservative 
minds will perhaps always prefer theories using this method of repre- 
sentation." 

''But the progress of physical research reveals facts that cannot be 
adequately covered by such intuitive theories. Progressive physicists 
therefore try to find new ways of representation, new theories which 
are adequate for as broad a domain of facts as possible, irrespective 
of whether the new methods supply the simplest representation of the 
facts that occur in everyday experience. It is thus possible that a new 
theory, although the simplest one thus far invented for handling sub- 
atomic phenomena, turns out to be quite complicated for representing 
such phenomena as apples falling from trees. For this reason such a 
theory is alleged to be incomprehensible to common sense, and con- 
servative minds are always reluctant to accept it. 

It is thus evident that the difference between the so-called two types 
of theory has little to do with the psychologic difference between those 
^physicists who like to reason abstractly and mathematically and those 
who like to use concrete images or models. If we examine carefully the 
work of modem physicists engaged in research on the quantum theory, 
we discover that the psychologic fact that some of them prefer mathe- 
matical formulas while others prefer geometric imagery does not ac- 
count for their attitude to the quantum theory. For the majority of 
physicists accept that theory, while only a minority denounce it, claim- 
ing that their “mental constitution” forces them to regard quantum 
theory as incomprehensible to common sense and as an unsuitable basis 
for laboratory research. Indeed, representatives of both psychologic 


151 



modern science and its phiiosophy 


types are found among those physicists who have contributed to the 
progress of quantum theory. 

The common impression that there is a profound gap between these 
two kinds of theory may perhaps be due to the fact that the funda- 
mental propositions of science are often formulated in what Carnap 
calls the “material idiom” instead of in the “formal idiom.” Thus, if we 
ask whether the physical world consists of bodies situated in space 
whose motions are governed by laws, or of purely mathematical prop- 
erties of a four-dimensional space-time continuum, we have formulated 
the question in the “material idiom” or “material mode.” So formulated, 
it leads to innumerable disputes of a metaphysical character. We can 
avoid such disputes by stating the question in a consistently scientific 
form, using the “formal idiom.” When so stated it becomes the follow- 
ing; Do the fundamental propositions of physics (i.e., its most general 
laws) contain only terms of the thing-language, or do they contain also 
terms that do not occur in the thing-language? ' It is important to ob- 
serve that even if the second alternative should be the case, the terms 
not occurring in the thing-language are reducible to terms in the thing- 
language; for otherwise the propositions containing them would be 
neither confirmable nor refutable in experience. It is clear, therefore, 
that when the question has been stated in this way, the difference 
between abstract and intuitive theories has ceased to have the funda- 
mental character often attributed to it. For when the formal idiom is 
used, the difference simply amounts to the difference between proposi- 
tions containing only thing-terms and propositions containing other 
than thing-terms which are, however, reducible to thing-terms. 

Since it is just our everyday experiences that are most conveniently 
formulated by the so-called intuitive theories, especially those using the 
language of mechanical models, the outstanding characteristic of such 
theories is that they give rise to the introduction of a thing-language. 

^ The term “thing-language” is used by Camap in the description of the world 
of our sense experience, as an alternative to “phenomenal” language. The latter 
decomposes our impressions of the physical world into elementary sensations, such 
as those of small areas of red or blue. The thing-language speaks of complexes of 
sense impressions as they occur in our everyday language, like “table,” "chair,” 
"man.” 


152 



modern physics and common sense 


And certainly an intellectual effort is required to overcome the reluc- 
tance to using a language that is not a thing-language, however useful 
such a new language might be for formulating the laws of physics about 
facts discovered by recent research. Accordingly, we can extract one 
grain of truth from the claim that intuitive theories (especially mechani- 
cal ones) are the only genuine physical theories that are compatible 
with the requirements of the laboratory scientist. This grain of truth lies 
in the postulate that every physical theory must contain only such 
terms as are reducible to terms of the thing-language. This postu- 
late, stated in other words, simply means that every physical theory 
must be tested in terms of propositions that deal with the facts of 
everyday experience. 

The theory of relativity has often been described as abstract, in 
contrast with theories about the ether, which are said to be intuitive or 
pictural. For the latter theories introduce a matter-like ether, which 
is asserted to behave in many ways like the bodies of everyday experi- 
ence, while relativity theory, on the other hand, introduces only ab- 
stract, disembodied formulas. Similarly, quantum mechanics, like rela- 
tivity theory, is said to introduce a system of purely mathematical 
formulas, which cannot be interpreted either in terms of waves, parti- 
cles, or any other visualizable things. 

Nevertheless, it is easy to see that the difference between classical 
mechanics and relativity mechanics has nothing to do with the sup- 
posed difference between a “natural” and an “artificial” or “sophisti- 
cated” theory. To characterize the difference between classical and 
relativistic mechanics in this way is as incorrect as a similar judgment 
on the difference between the geocentric and heliocentric theories in 
astronomy. 

Classical mechanics is very convenient for describing and pre- 
dicting motions of bodies with velocities small in comparison with the 
velocity of light. Such comparatively slow motions include the move- 
ments of bodies we observe dmly, even the movements of the celestial 
bodies such as the sun, moon, and stars. In order to describe these mo- 
tions it is convenient to introduce the term “length of a body” without 


153 



modern science and its philosophy 


specifying any frame of reference with respect to which this length 
is to be measured. Similarly, we use tlie expression “two events occur 
simultaneously” without requiring the specification of a frame of ref- 
erence in order that the expression should have a definite meaning. 

However, the situation becomes altered if we try to describe and 
predict the motions of bodies having velocities comparable with the 
velocity of light, for example, the fi-iays of radium, or cathode rays. 
In order to formulate in the simplest way the laws governing such 
rapid motions, it is found most convenient to drop the term “length of 
a body” and introduce the new term “length of a body A with respect 
to a body B“ where B has a determinate velocity with respect to A. 
The definition of the “relative length of a body” will consist in the 
description of the procedures of measurement involved, just as in the 
case of the definition for the so-called “absolute length” of a body. 
Since all these procedures of measurement are described in the thing- 
language of daily life, there is no occasion to regard one of these defini- 
tions as more abstract than the other. The definition of “relative 
length” involves the use of observation terms just as does the definition 
of “absolute length,” the sole difference being that the first definition 
is slightly more complicated. The use of “relative length” in the de- 
scription of the motions of bodies occurring in daily experience would 
complicate those descriptions unnecessarily; but it would not make 
them either too abstract or unintuitive. On the other hand, were we 
to use the term “absolute length of a body” in describing the behavior 
of bodies having velocities comparable with that of light, we would 
also greatly complicate our descriptions; that is to say, our system 
of physics dealing with such bodies would become very inconvenient if 
we persisted in using the term ‘length of a body” according to the 
formative rules by which the syntax of the language of classical physics 
is determined. The most familiar and popular example of such a com- 
plication is the necessity of introducing the terms “ether” and “ve- 
locity with respect to the ether” into the system of physics. These terms, 
however, remain in a sense isolated from most other terms of physics, 
simply because no experiment can be set up by which we could deter- 
mine the magnitude of this “velocity with respect to the ether.” 


154 



modern physics and common sense 


Einstein’s theory of relativity in effect introduces new formative 
rules into the system of physics for the use of the term “length of a 
body.” According to these rules the term “length of a body” can occur 
in sentences only if the sentences also contain such qualifying expres- 
sions as “with respect to a specified reference frame.” In this way super- 
fluous terms like “velocity with respect to the ether” can be eliminated 
from the language of physics, while at the same time the laws govern- 
ing high velocities like those possessed by j8-rays can be formulated 
in a very simple way. 

S imilar considerations hold for the term “simultaneously.” In 
classical mechanics it is used in accordance with a certain set of forma- 
tive rules which do not require that sentences containing the term also 
contain the qualifying phrase “with respect to a specified frame of 
reference.” In relativistic mechanics we employ a different set of 
formative rules; according to these rules the term “simultaneously” 
may be employed only if the indicated qualifying phrase also 
occurs. 

These syntactical differences between classical and relativistic 
mechanics are often expressed in ways tliat easily lead to misunder- 
standings. Thus, it is sometimes said that classical mechanics con- 
siders simultaneity in an “absolute sense” while relativistic mechanics 
does so in a “relative” sense. It is advisable, however, to state the 
matter intended by using the “formal idiom,” and to express the 
difference in terms of the different formative rules according to which 
the word “simultaneous” is used in classical and relativistic mechanics. 
For if we talk of “simultaneity in an absolute and in a relative sense” 
we are easily led to the meaningless question whether absolute simul- 
taneity exists or not. 

It is clear that there is no difference between these two sets of for- 
mative rules in their being “abstract” or “intuitive” on the one hand, 
or “natural” or “intuitive” on the other. The sole significant difference 
between them is that the language of classical mechanics together 
with its formative rules is more convenient for discussing motions of 
bodies that occur in daily experience, while the language of relativistic 
mechanics and its distinctive formative rules is more suitable for for- 


155 



modern science and its philosophy 


mulating the more inclusive field of phenomena which contains bodies 
with high velocities. 

Anyone who supposes that the formative rules of relativistic syntax 
are artificial, nonintuitive, or abstract in comparison with the rules for 
classical mechanics can readily convince himself of the contrary by 
imagining himself in a world in which the phenomena of daily experi- 
ence occur in a way somewhat different from the way they usually do. 
It will then be evident that both in the case of quantum physics and 
in the case of relativity theory, the laws alleged to be in accordance 
with “common sense” will be of quite a different type from what is 
usually maintained. 

Let us, therefore, in the manner of H. G. Wells, imagine a world in 
which the flight of a tennis ball ca nn ot be formulated and predicted in 
terms of Newton’s laws. Thus we may suppose that the path taken 
by a ball is not determined by the way in which it is struck with the 
racket. We also suppose that the only law which can be established in 
this world simply states with what percentage balls struck in a certain 
way take a determined direction with a determinate velocity. Tennis 
matches could be organized in this world just as well as in the ordinary 
Newtonian world, and skilled players would return served balls more 
often than those not skilled. Although the game would be played in 
accordance with the ordinary rules of tennis, it would be evident to a 
careful observer that even the most proficient players would often not 
succeed in sending the ball in an intended direction even when the ball 
is struck in the appropriate way. 

In such a world neither the familiar man-in-the-street with his 
well-known common sense, nor the laboratory physicists with a critical 
attitude toward all “abstract” theories, would claim it to be a matter 
of common sense that the laws governing the motions of tennis balls 
should be formulated in terms of the expression “the position and 
velocity of a tennis baU at the moment after it is struck.” For such an 
expression would not occur in any of the laws with the help of which 
the direction of flight of a struck ball could be predicted. Indeed, in this 
imagined world the laws would require a specification of the way in 


156 



modern physics and common sense 


which a ball is struck, but would have no use for the specification of the 
velocity a ball acquires on being hit. In brief, even everyday experi- 
ence in this world would not be formulated in laws containing the 
expression “ball with a determinate position and velocity.” The ex- 
pression would not be employed in the language of daily affairs, and 
would be taken to be as meaningless for everyday experiences as it in 
fact is in the actual world of atomic physics. For in this imagined 
world tennis balls behave in the fashion of the atoms and their nuclei 
of our actual world. 

The impression often current that the language of quantum physics 
is artificial and contradictory to common sense arises from the fact 
that this language is unnecessarily exact for use in the affairs of daily 
experience. If, however, the things in daily experience behaved as ten- 
nis balls would behave in our imagined world, this language would 
be the one most suitable for discussing even the most familiar mat- 
ters. In that case no one would dream of claiming that the language of 
quantum physics is artificial or that it is incompatible with common 
sense. It would be accepted as a very natural language indeed, and 
would be judged to be as convenient and suitable for everyday mat- 
ters as quantum mechanics is now taken to be for the phenomena of 
atomic nuclei. 


157 



CHAPTER 


8 


philosophic misinterpretations of the quantum theory 

1. The Origin of Philoiophie interpretations of Physicol Theories 

AS soon as any new physical theory appears, it is used to con- 
tribute something toward setting the controversial questions 
of philosophy, the questions on which philosophers have 
been working for centuries without coming a single step closer to 
their solution. Numerous examples of such use suggest themselves. 
When J. J. Thomson showed that every electrically charged particle 
possesses inertia just as a mechanical mass does and gave a formula 
for calculating the mechanical mass of a particle from its charge and 
size, people deduced from this result arguments to prove that all 
matter is only a phantom. They found in it an argument for the ideal- 
istic world view and against materialism. Similar interpretations were 
suggested when energetics arose and phenomena were represented as 
energy transformations rather than as arising from collisions of masses. 
The theory of relativity then introduced four-dimensional non- 
Euclidean space instead of the three-dimensional Euclidean space in 
which the directly observable processes of everyday life take place. 
Later, wave mechanics described physical processes with the aid of 
the probability concept, which has been often said to be a purely 
spiritual factor, instead of with the aid of mass particles. Everywhere 
it appears that the spiritual element is replacing the grossly material. 

Such interpretations were attached with especial intensity to Niels 


158 



ghilosophie misinterpretations of quantum theory 

Bohr’s theory of the complementary nature of certain physical descrip- 
tions, from which it was hoped that arguments for vitalistic biology 
and for free will might be obtained. 

If one scans all such interpretations, the empirically establishable 
fact is found that they all further a movement toward a certain world 
picture. It is not a case of different world pictures being involved; the 
same one keeps coming up again and again. 

Through the work of Galileo and Newton, anthropomorphic medi- 
eval physics was expelled from conscious intellectual life. There re- 
mained, however, an unfulfilled longing to bring about the unity of 
animate and inanimate nature which had been present in medieval 
physics but was missing in the newer physics. There was left only one 
problem, for which no satisfactory solution could be envisaged: to 
understand the processes of life in terms of physics. For that was the 
necessary condition for a unified conception of nature after the dis- 
appearance of the anthropomorphic conception of physics, which had 
fitted in so well with the vitalistic conception of life. 

Every crisis in the history of physical theories is associated with a 
certain lack of clarity in their formulation, and this unfulfilled longing 
bursts forth with great strength from the unconscious. Efforts were 
made to complete the new physical theories by "philosophic interpreta- 
tions’’ in order to proclaim the imminence of a return to the anthro- 
pomorphic physics of the Middle Ages and a consequent reestablish- 
ment of the lost unity of nature. Spiritualistic physics was to hold the 
possibility of embracing the living processes also. 

The assertion is often heard that there is also a philosophic interpre- 
tation of physical theories in the service of a materialistic-metaphysical 
picture of the world. However, this symmetrical conception of spiritual- 
ism and materialism is a very superficial one. A “materialistic meta- 
physics,’’ in general, does not exist today as a living intellectual current. 
At most, it is represented for the domain of physics by those philoso- 
phers or scientists who wish to make as wide a gulf as possible between 
physics and biology, in order to obtain in the field of the living or the 
social processes free play for a spiritualistic metaphysics. 

If, on the other hand, one understands by materialism the belief 


15? 



modern science and its philosophy 


that all processes of nature can be reduced to the laws of Newtonian 
mechanics, then this is not a philosophic principle but a physical 
hypothesis. True, it is a physical hypothesis that has been shown to be 
wrong, but a physical statement it remains. This false hypothesis is not 
accepted today by any of the philosophic schools that one is accus- 
tomed to designate polemically as “materialistic"— neither by the 
“dialectical materialism” of Soviet Russia, nor by the “physicalists” 
that have come out of the Vienna Circle. 

The process of the philosophic interpretation of physical theories in 
the service of the spiritualistic conception of the universe can be an- 
alyzed both psychologically and logically. From the psychologic stand- 
point, the following, roughly, has been established: The physicist, like 
every other educated person, acquires the remnants of prescientific 
theories as a “philosophic” world picture, which in our cultural circles 
consists mostly in a vague idealism or spiritualism as it is usually 
learned from lectures on general philosophy. The principles of this 
philosophy are unclear and difficult to understand. The physicist is 
happy if he finds in his science any propositions that have in their 
formulation some similarity to propositions of idealistic philosophy. 
He is often very proud that his field of work helps him throw some 
light on the general doctrines that are so important for this world 
picture. Thus even the slightest similarity in the wording is enough 
to induce the physicist to offer a proposition of his science as support 
for the idealistic philosophy. 

If J. J. Thomson speaks of “real” and “apparent” mass, the philo- 
sophically educated physicist is eager to bring this mode of expression 
into connection with the distinction between a “real” and an “apparent” 
world. The statement that mechanical mass is only "apparent” mass is 
then taken as confirmation of philosophic idealism, according to which 
matter is only an illusion. 

Of greater scientific interest is the logical structure of these philo- 
sophic misinterpretations. The process of thought leading to them 
consists of two steps. First, physical propositions that are really state- 
ments about observable processes are regarded as statements about 
a real, metaphysical world. Such statements are meaningless from the 


160 



philosophic misinterpretations of quantum theory 

standpoint of science, since they can be neither confirmed nor con- 
tradicted by any observation. The first step is therefore the transition 
to a meaningless metaphysical proposition. In the second step this 
proposition, by means of a rather small change in wording, goes over 
into a proposition which again has a meaning, but is no longer in the 
realm of physics; it now expresses a wish that people should behave 
in a certain way. This proposition is then no longer metaphysical, but 
has become a principle of morality, of ethics, or of some other system 
of conduct. 

One can adduce numerous examples of such processes involving 
two steps. As the simplest, we choose the well-known example of the 
electromagnetic mass. J. J. Thomson formulated the purely physical 
proposition that every electrically charged body possesses mechanical 
inertia, which can be calculated from the charge. To this has been 
added the hypothesis, likewise physical, that the entire mass of the 
body can be calculated in this way. Philosophers then expressed this 
as a metaphysical principle by saying: “In the real world there is no 
mechanical mass at all.” This principle obviously has no scientific con- 
tent. From it there follow no observable facts. As the second step, it 
was asserted that the material world, as a mere illusion, is unimportant 
in comparison to the world of the spirit, and that therefore man in his 
actions can or should neglect any changes in the material world and 
should devote himself to his spiritual perfection. 

When influential groups express such wishes, the fact has a great 
importance for human life, of course, and possesses a meaning, but 
there evidently exists no logical connection with the electromagnetic 
theory of matter, and the whole thing arises only through this mis- 
interpretation with its two steps. 

The essential part of the misinterpretation is the passage through 
the “real” metaphysical world. The misinterpretation can therefore 
be avoided only if one tries to set up a direct short circuit between 
the physical principle and the moral principle. This can be done, for 
example, through the consistent use of the “physicalistic language,” 
which Neurath and Carnap have suggested as the universal language 
of science. 


161 



modern science and its philosophy 


As understood in Carnap’s “logical syntax,” the source of these 
misinterpretations is always the use of the “material mode of speech.” 
The contrast between “apparent” and “real” mass is made to appear 
as a statement about a fact of the observable world, whereas it is really 
a syntactic rule about the use of the word “real.” Only the formula for 
the connection between electric charge and inertia is a statement about 
the observable world. 

Quite the same logical structure is possessed by the misinterpreta- 
tions of the relativity and quantum theories. The first has been em- 
ployed to provide a basis for the belief in predestination, the second 
to give scientific arguments for “spontaneity of action” and “freedom of 
the will.” 


2. The Complementary Conceptions of Quantum Mechanics and Their Interpretotions 

The philosophic misinterpretations of quantum mechanics can be 
best understood if we remember that the same tendencies are at work 
here as in the interpretation of previous theories, and that the process 
takes place along exactly the same lines, both psychologically and 
logically. 

First we must make clear the meaning of the complementarity 
conception in physics. 

One often reads the following formulation: “It is impossible to 
measure the position and the velocity of a moving particle simul- 
taneously.” The world, therefore, just as it is according to classical 
mechanics, is filled with particles having definite positions and veloci- 
ties; unfortunately, we can never attain a knowledge of them. This 
presentation, in which the states of the particles play the role of the 
“thing in itself” in idealistic philosophy, leads to innumerable pseudo 
problems. It introduces physical objects, namely, particles with definite 
positions and velocities, about which the physical laws of quantum 
mechanics say nothing at all. These objects play a role similar to that 
of the reference system that is absolutely at rest, which some wish to 
add to the theory of relativity but which never occurs in any physical 
proposition. In both cases the reason for this addition is that such 
expressions were found useful in the earlier state of physics, and the 



philosophic misinterpretations of quantum theory 

school philosophy had made of them constituents of the “real world”; 
therefore they must be kept forever. 

Another way of representing the situation consists in saying that 
particles “in general do not possess definite positions and velocities 
simultaneously.” This mode of expression appears to me to have the 
diflBculty that the combination of words “particle with an indefinite 
position or velocity” transgresses the syntactic rules according to which 
the words “particle,” “position,” and “indefinite” are ordinarily used 
in physics and everyday life. Of course, there would be no objection 
if a new syntax were introduced for these words for the purposes of 
quantum mechanics. In that case, expressions like “particle with an 
indefinite position” could be employed inside of physics without any 
danger. And there exist many correct works on the quantum theory in 
which this is the case. However, gross misunderstandings arise as soon 
as this way of speaking is used in matters where it is no longer a ques- 
tion of the quantum theory. We can bring about this transition to 
other fields only by regarding the particle with an indefinite position 
as a constituent of the “real world”— and then we are right in the 
midst of the philosophic misinterpretations that were described in 
Section I. 

I believe that, as a starting point for a correct formulation of the 
complementarity idea, one must retain as exactly as possible the 
formulation set forth by Bohr in 1936. 

Quantum mechanics speaks neither of particles the positions and 
velocities of which exist but cannot be accurately observed, nor of 
particles with indefinite positions and velocities. Rather, it speaks of 
experimental arrangements in the description of which the expressions 
“position of a particle” and “velocity of a particle” can never be em- 
ployed simultaneously. If in the description of an experimental arrange- 
ment the expression “position of a particle” can be used, then in the 
description of the same arrangement the expression “velocity of a 
particle” can not be used, and vice versa. Experimental arrangements, 
one of which can be described with the help of the expression “position 
of a particle” and the other with the help of the expression “velocity” 
or, more exactly, “momentum,” are called complementary arrange- 


163 



modern science and its philosophy 

ments, and the descriptions are referred to as complementary descrip- 
tions. 

If one adheres strictly to this terminology one will never run the 
risk of falling into a metaphysical conception of physical comple- 
mentarity. For it is clear that nothing is said here about a “real world,” 
nor about its constitution, nor about its cognizability, nor even about 
its indefiniteness. 

A great seduction to metaphysical interpretations lies in the fre- 
quently occurring formulation of complementarity according to which 
the “space-time” and the “causal” descriptions are said to be comple- 
mentary. In this way the fact is often hidden that this again only means 
the complementarity of position and momentum, or of time and energy. 
By “causal description” we understand here only the description by 
means of the principles of conservation of energy and of momentum, 
which does not quite agree with what is usually understood by 
causality. In popular presentations, among which are those of some 
physicists, this is not always set forth clearly. This lack of clarity arises 
from the use of the expressions “space,” “time,” and “causality,” which 
as a kind of trinity play a somewhat mysterious role in idealistic phi- 
losophy. If by “space-time description” is meant simply the assignment 
of coordinates and time, by “causal description” the application of the 
conservation principles, then this beloved terminology can be retained, 
of course. But it then loses the charm of the mysterious and can no 
longer be used to pave the way for a transition from physics to ideal- 
istic philosophy, thereby favoring those misunderstandings described 
in Section I. 

If we are once in the midst of metaphysical formulations, we can 
easily come to rather crass misinterpretations. As an example, I shall 
give one by a very prominent physicist. A. Sommerfeld says ( Scientia, 
1936 ): 

If we treat the human body physiologically, we must speak of a corpus- 
cular localized event. To the psychic principle we can assign no localization, 
but must treat it— and this is also the opinion of psycho-physiologists— as if it 
were present more or less throughout the body, just as the wave is connected 
with the corpuscle in an unspecifiable way. 


164 



philosophic misinterpretations of quantum theory 


Here we may see with great clarity how every metaphysical formula- 
tion of a statement of physics can be used with great ease to support 
a statement of idealistic philosophy that only sounds somewhat similar. 

To express the idea of complementarity for physics in closest as- 
sociation to Bohr’s formulation, so that it will not lead to any meta- 
physical misinterpretations but yet can be carried over to fields out- 
side of physics, one will have to proceed somewhat as follows: 

The language in which occur statements like “The particle is at this 
place and has this velocity” is suited to experiences involving gross 
mechanical processes and cannot be employed satisfactorily for the 
description of atomic processes. However, one can give a group of 
experimental arrangements for the atomic domain in the description 
of which the expression “position of a particle” can be used. In the 
description of these experiments— and in this consists the idea of Bohr 
—the expression “velocity of a particle” can not be used. In the atomic 
domain, therefore, certain parts of the language of gross mechanics 
can be used. The experimental arrangements, however, in the descrip- 
tion of which these parts can be used, exclude each other. 

Meaningless metaphysical propositions immediately arise if one 
says that "reality” itself is “dual” or displays “different aspects.” 


3. Complementarity as an Argument for Vitalism and Free Will 

Many physicists and philosophers have tried to make use of Bohr’s 
doctrine of the complementarity of physical concepts in order to ob- 
tain arguments for the impossibility of an understanding of biology and 
psychology in terms of physics. Here we can distinguish something 
like a psychologic and a biologic argument. The first runs approximately 
as follows: If one seeks to describe a psychic state in terms of intro- 
spective psychology, the state is so strongly altered by self-observation 
that it is no longer the original state. It is not possible to be angry and 
at the same time to observe and describe one’s anger. The existence of 
a psychic state is incompatible with its observation. 

The second runs something like this: If one wishes to describe the 
state of a living organism by means of physical quantities, the measure- 
ment of these quantities requires such a severe disturbance of the 


165 



modern science and its philosophy 


organism that it must be killed. The description of a living being 
through physical variables is incompatible with its life. 

The psychologic argument is basically a good one. It is a long- 
recognized doctrine of every positivistic conception of science, in- 
cluding that of A. Comte, that one cannot found any logically con- 
nected psychology on principles obtained through self-observation. 
One must go over to an objective observation of human actions and 
movements of expression, as required by American behaviorism, and in 
accordance with the logical analysis given by Camap and Neurath of 
the statements concerning psychic processes. 

If psychology is formulated in terms of behaviorism or physicalism, 
the psychologic argument coincides with the biologic one. 

If one applies the Bohr idea of complementarity, one can formulate 
the role of self-observation in psychology somewhat as follows: There 
are certain experimental arrangements in the field of psychology that 
can be described with the aid of propositions and expressions obtained 
from self-observation. There are other situations in our life that cannot 
be described with these expressions. In this there is no contradiction. 
As in physics, so in psychic life there are complementary situations, and 
complementary languages for their description. 

Taking this complementarity into consideration, one will easily see 
what can be gained for the understanding of free will from the analogy 
to the quantum theory. Even before Bohr’s discovery of complemen- 
tarity, M. Planck had advanced the following argument for the com- 
patibility of free will with physical causality: If a man could calculate 
his future actions from the present pattern of the physical world, this 
knowledge would react on his present state— for example, on the 
molecules of his brain— and thus change his state. Hence there is no 
predictability of the future. Hence free will cannot be in contradiction 
to the physical causality of occurrences in the human body. 

From this it only follows that a man cannot calculate his future 
actions from the results of self-observation. It might still be possible, 
however, for one to calculate beforehand the actions of other men, and 
to do so even from purely physical observations. 

If one applies here Bohr’s idea of complementarity, one can give 


166 



^philosophic misinterpretations of quantum theory 

to the whole matter a firmer logical structure. One can then say: Cer- 
tain situations of human behavior are described with the help of the 
expression “free will”; under other experimental conditions this ex- 
pression cannot be used. We are, therefore, dealing here with comple- 
mentary situations and with complementary descriptions, but not 
with any contradictions. Bohr himself pointed out that his considerations 
of complementarity cannot be used to provide an argument for “free 
will”; they can only yield a useful representation of the epistemologic 
status of the problem. 

It seems to me, however, that there is also a certain objection to 
the use of the words “free will” for the description of certain situations, 
corresponding to the experimental arrangements in physics. Expressions 
like “position of a particle” are expressions taken from the physics of 
everyday life which, because of complementarity, remain suitable 
for atomic physics only in certain special situations. Likewise, “free 
will” would have to be an expression from the psychology of daily 
life which in scientific psychology could be employed only under 
certain experimental conditions. This, however, seems to me not to be 
the case. “Free will” is not an expression from the psychology of daily 
life; it is rather a metaphysical or theological expression. In everyday 
life “freedom” is never anything other than “freedom from external 
coercion,” or at most “freedom from intoxication and hypnosis.” This 
has nothing to do with the philosophic conception of freedom of will. 
If it is correct to say, following Bohr, that the expression “free will” 
can be used advantageously for the description of certain situations, 
this expression can refer only to the quite unphilosophic concept drawn 
from the psychology of everyday life. Hence from this use no con- 
clusions can be drawn about the philosophic freedom of will. It is only 
necessary to put to oneself the question whether, for the general situa- 
tion in which the concept of free will is used in practice, any change has 
been created by quantum mechanics and the complementarity con- 
cept. By this I mean, of course, the application of the freedom concept 
to the question of the responsibility of a criminal, and to the related 
question of the harshness or lightness of the punishment. One need only 
formulate precisely die whole idea of complementarity and follow 


167 



modern science and its philosophy 


through carefully the whole chain of ideas up to the punishment of 
the criminal to see at once that no consequences follow here for the 
problem under consideration. It is therefore very questionable whether 
it is appropriate to use the expression “free will” in the applications of 
the complementarity idea to psychology. 

If, however, in accordance with the new conceptions of behavior- 
ism and physicalism, psychology is based on principles containing, 
not statements about self-observation, but statements about the be- 
havior of experimental subjects, then the complementarity considera- 
tions in psychology as just described drop out, and psychology be- 
comes a part of biology. In that case the psychologic argument of 
Bohr reduces to the biologic one. It is, therefore, a question of whether 
the behavior of living organisms can be represented by laws in which 
only physical variables occur. 

If one wishes to describe a living being physically, one must specify 
the state of each of its atoms: this is Bohr’s starting point. The observa- 
tions required for this description, however, involve physical disturb- 
ances of the organism that are so great as to be fatal. The states of the 
atoms of an inanimate body can be specified within the limits imposed 
by the Heisenberg uncertainty relations, whereas the large protein 
molecules with which life is associated are destroyed by disturbances 
that would allow atoms to continue to exist. 

Experiments by which the living organism may be described in 
terms of the functions that characterize it as living are therefore carried 
out under experimental conditions quite different from those of 
experiments on the organism as a physical system. According to Bohr, 
it is a question here of “complementary” e,\perimental arrangements, 
which are described in “complementary languages.” Therefore, to 
describe the phenomena of life in a language which is not that of 
’ physics or chemistry is logically free from objections and does not 
constitute a lapse into a spiritualistic vitalism. 

This way of putting the matter, as given by Bohr, is very different 
from that of most of his "philosophic interpreters,” and is certainly 
tenable. In so far as its usefulness is concerned, some remarks can be 
made. The whole argument derives its force from the fact that it is an 


168 



philosophic misinterpretations of quantum theory 


analogue of the argument that led from classical to quantum physics 
and justified the statement that atomic processes cannot be described 
in the language of classical physics. In order to establish limits for 
the appropriateness of this analogy we shall therefore compare two 
lines of thought. 

First, in the transition to quantum physics one reasons as follows: 
According to classical physics, one must be able, in principle, to devise 
experiments permitting the measurement of the positions and the 
velocities of individual particles veith arbitrary accuracy. But our 
knowledge of atomic processes— for example, the Compton effect- 
shows on closer analysis that the possibility of such measurements is 
contradicted by experience. Hence atomic phenomena cannot be 
described in the language of classical physics. 

If we wish to extend this reasoning from the inanimate to the 
animate, we must accept it as an experimental fact that an observation 
by physical means, sufiSciently accurate to enable one to describe 
exactly the physical state of the individual atoms of a living body, 
represents so great a disturbance that it kills the organism. It follows 
then that classical physics, aided by quantum physics (of inanimate 
atoms), is inadequate for the description of the phenomena of life, 
since it is incompatible with the application of physics to the living 
organism that the latter should be killed by every act of exact 
measurement. 

The strength of the quantum theory lies in the fact that no hypoth- 
esis about the atom based on classical physics could be found that 
was in agreement with the experimentally testable behavior of observ- 
able bodies. If the testing of a hypothesis about atoms through direct 
measurements of their mechanical state (position and velocity) had 
not been in contradiction with empirical facts, the hypothesis would 
have remained within the framework of classical physics. Since, how- 
ever, quantum mechanics does involve contradiction, it goes beyond 
classical physics. 

If we wish to retain the same chain of ideas for the transition from 
inanimate to living bodies, then empirical evidence must be presented 
to show that the exact physical observation of the atoms of a living 


169 



modern science and its philosophy 


^ body is incompatible with the known empirical laws for the behavior 
of living bodies and with the physical hypothesis about tlieir atomistic 
structure. As long as this evidence has not been submitted, it follows 
only from Bohr’s train of thought that in biology, in the present state 
of our knowledge, the complementarity mode of expression is possible 
and perhaps even desirable. In contrast, for the transition from classi- 
cal physics to quantum mechanics one can conclude that in atomic 
physics the complementarity mode of expression is necessary. 


4. Summorizing Remarks 

From all that has been said, it is clear that Bohr’s complementarity 
theory does not provide any argument for free will or vitalism. Like- 
wise, one cannot derive from it any new conception about the relation 
between the physical object and the observing subject, if we under- 
stand the words “object” and “subject” in the sense in which they are 
used in empirical psychology. In presentations of quantum mechanics 
in which reference is made to this new role of the observing subject, 
the word “subject” is understood in quite another sense. By “subject” is 
always meant the measuring arrangement, which can be described in 
terms of classical physics. What was shifted by the quantum theory 
was the relation between the object of atomic theory— the atom or 
electron— which cannot be described by means of classical physics, and 
the measuring instrument, which can be described classically. The 
observing subject, in the sense of empirical psychology, has no other 
task than to read off the measuring instrument. 'The interaction be- 
tween measuring instrument and observing subject can be described 
classically, as far as we can say from the present state of physics. 
The boundary line between the classical and the quantum-mechanical 
descriptions lies between the electron and the measuring instrument. 
Since within the region of classical description it can be displaced 
arbitrarily, the boundary line can also be drawn between the measur- 
ing instrument and the observer. But thereby nothing new is expressed, 
since within the classical region the position of the section is arbitrary. 

The great importance of Bohr’s complementarity theory for all 
branches of science, especially for the logic of science, seems to me 


170 



philosophic misinterpretations of quantum theory 

that it starts out with a language that is generally understood and 
accepted, the language used to describe the gross mechanical processes 
of motion. Its significance lies in the fact that in its use all men are in 
harmony. In physics this language is used in such expressions as “posi- 
tion of a particle,” in the sense of gross mechanics. Atomic processes, 
however, cannot be described in this language, as the new physics has 
shown. Bohr has demonstrated in a careful analysis of modem physics 
.that certain parts of the language of everyday life can nevertheless be 
retained for certain experimental arrangements in the field of atomic 
phenomena, although different parts are required for different experi- 
mental arrangements. The language of daily life thus possesses comple- 
mentary constituents which can be employed in the description of 
complementary experimental arrangements. 

There is no doubt that this idea is also a fruitful one for logical 
syntax in general and deserves to be applied to other branches of 
science. One would have to start out in psychology with the language 
of everyday life and see whether, in the transition to more subtle prob- 
lems, this language could be retained. One might perhaps start with 
the “physicalistic” "protocol language” of Camap and Neurath and see 
whether any parts of it are particularly suitable for describing certain 
situations. Perhaps the symbol language of psychoanalysis is a sugges- 
tion of such a partial language. The phenomenal language of which 
Camap often spoke in his earlier works must be dropped as a general 
language, but perhaps, as a constituent of a general language in the 
sense of the Bohr conception, it can provide a satisfactory description 
for certain experimental situations. 


171 



CHAPTER 


9 


determinism and indeterminism in modern physics 

% A #HEN philosophers describe their standpoint with regard 
\l\Mto a ne w physical the ory, their attitude is usually one of 
Y f the following three; (1) the new theory contradicts the cor- 
rect philosophic system and is therefore false; (2) the new theory 
is a brilliant confirmation of the correct philosophic system and is there- 
fore to be welcomed; (3) the new theory can be used for more or less 
important improvements in the correct philosophic system and there- 
fore possesses a certain value. 

When the physicist attempts to continue outside of physics his 
manner of forming principles and of verifying them through experi- 
ments, he comes to the conception of science called l ogical empirici sm. 
This conception has become fairly well known in recent years, es- 
pecially through the work of the Vienna Circle. It is from this point of 
view that I wish to examine a book ^ by Ernst Cassirer, since this is 
the point of view most closely associated with the thinking of the 
physicist. 

No doubt many will demand a so-called “immanent criticism” of 
a philosophic work. I believe, however, that such a demand usually 
signifies only a plea of extenuating circumstances, and in the case of 

^E. Cassirer, Determinismus und Indeterminismus in der modemen Phystk; 
historische und systematische Studien zum Kausalproblem (Goteborg; Elanders 
Bokhyckeri Aktiebolag, 1937). 


172 



determinism and indeterminism in modern physics 

a book as important as the present one this would be even derogatory. 
For, at best, the result of immanent criticism of a book is the ac- 
knowledgment that even if it is nonsense, it does have method. 

Such rather scholastic criticism I will not attempt. Instead I want 
to consider how one must judge Cassirer’s exposition from the stand- 
point of logical empiricism, according to which only those statements 
may occur in science that can be justified through logical derivation 
or empirical tests. 

According to this conception, philosophic principles, which are not 
scientific in the above-mentioned sense, form a system of isolated 
propositions from which there are no logical bridges to the system of 
scientific propositions. Hence a system of philosophic principles can 
never be either confirmed or refuted by new physical theories. Strictly 
speaking, neither can it undergo any improvement. There is often 
the appearance of improvement, but it can arise only from the fact 
that an agreement in emotional coloring is taken for a lo^cal agree- 
ment. 'This is frequently made possible because the physical principles 
are formulated in a metaphysical, rather than a purely physical, lan- 
guage. In such a case, however, one should not say, for example, that a 
physical theory is in contradiction to a philosophic system, but rather 
that the metaphysical formulation of the theory appears to be irrecon- 
cilable with the philosophic principles under discussion. 

When one reads this book by Cassirer one gets at once the impres- 
sion that the foregoing considerations about the relation between sci- 
entific and philosophic principles do not apply to it. Most of Cassirer s 
assertions about the new physical theories are acceptable throughout 
from the standpoint of the physicist and its extension into logical em- 
piricism. I might say even more. Cassirer’s statements about the new 
quantum mechanics are less full of metaphysical prejudices and con- 
tain fewer transitions from physical to metaphysical language than 
some statements made by physicists in their professional writings or 
in speeches on festive occasions. Cassirer’s statements are almost all 
scientific statements, as understood by logical empiricism. Isolated 
systems of propositions, such as those that play the chief role in the 
school philosophy, hardly occur. Hence there cannot arise any appar- 


173 



modern science and its phiiosophy 


ent contradiction between physical and philosophic principles. What 
does induce in me a critical state of mind is a certain background 
against which the presentation is brought into relief, a background 
which is distinctly separated from the basic assertions, but which 
through its terminology, foreign to science, makes me feel again and 
again that the author regards all that he says as only rather superficial 
utterance, the deeper meaning of which he intimates but does not 
wish to discuss. 

For this reason I have designated Cassirer’s mode of thinking, 
which is to be seen not only in his present writing but also in his earlier 
works, a "disintegration process within the school philosophy.” 

This remark in my book “The Law of Causality and its Limits” “ 
has often been looked upon as an unfavorable criticism of Cassirer’s 
point of view. Quite the opposite is the case. In the book mentioned I 
wanted to show that the “disintegration” of the school philosophy 
forms a necessary preliminary condition for the progress of science 
from the individual sciences to a unified science. Precisely in the case 
of the present book it can be made clear what the disintegration process 
that Cassirer carries out consists in, and, furthermore, we can also see 
how this process remains within the school philosophy, at least in its 
emotional background. Practically considered, Cassirer’s conception 
of the principles of physics is almost exactly that of logical empiricism. 
However, at the end the sharply drawn contours are a little blurred in 
the direction of transcendental idealism— which is perhaps only a 
question of style. It seems to me that this shading of the outlines is a 
little dangerous, since we know that the reader is often influenced more 
by the emotional undertone of an exposition than by its logical and 
empirical content. 

Some of the basic conceptions of Cassirer are in a remarkable 
manner related to, or almost coincident with, the conceptions formed 
by the logical empiricism of science. Cassirer repeatedly points out 
that science creates auxiliary concepts, such as force or atom, in order 
to be able to formulate conveniently the theories it has set up at a 

*P. Frank, Das Kausalgesetz und seine Grenzen (Vienna: J. Springer, 1932); 
La Loi de causalitS et ses Unites (Paris: Flanunarion, 1936). 


174 



d[,eterininism and indeterminism in modern physics 

certain time, but that at later times these auxiliary concepts freeze into 
essences, “ontological concepts,” which are retained even if they are 
no longer very convenient in the state of science at that time. This, 
however, is precisely the criticism leveled at the existence concepts of 
the philosophers by the Vienna Circle, so often attacked as “anti- 
philosophic.” 

In my essay on the death of Ernst Mach,* I characterized as the 
enduring nucleus of Mach’s teachings his struggle against the “idoliza- 
tion of auxiliary concepts.” The French Catholic philosopher J. Mari- 
tain, at the Thomistic Congress in Rome in the summer of 1936, 
characterized as a great service that was essential also for Catholic 
philosophy the fact that the aim of the Vienna Circle and of the whole 
movement of logical empiricism was "to disontologize scienc e." 

The disintegration of the school philosophy in Cassirer’s work is 
revealed very clearly in his general conception of the law of causality. 
He believes that there is no such law that can be formulated like a 
definite law of nature. In his opinion the law of causality asserts only 
that there exist, in general, laws of some sort in nature. Kant tried to 
formulate causality also with respect to its content, by asserting, for 
instance, that for every process there exists another “which it follows 
according to a rule.” Cassirer rejects not only the very special world 
formula of Laplace, but also the far more general formulation of Kant; 
there remains only the formal requirement, also attributable to Kant, 
that nature be describable with the help of simple rules. Here, how- 
ever, there is left hardly any contradiction to Ernst Mach’s purely 
positivistic conception of science. 

Another fundamental feature of Cassirer’s conception is that the 
form of the law of causality and the concepts of what one calls an 
“object” mutually condition each other. This also is a basic thesis of 
logical empiricism, and one that has been taken over from positivism. 
It is only that logical empiricism gives to this thesis a more formal 
turn: Ae form of every physical law depends on what variables are 

*P. Frank, "Die Bedeutung der physikalischen Erkenntnistheoiie Machs fur 
das Geistesleben der Gegenwart,” Natutwisseruchaften 5, 65 (1917). This essay 
is Ghapter 2 of the present hook. 


175 



modern science ond its philosophy 


introduced to describe the state of a system. It may happen that with 
respect to some variables for a certain domain of phenomena there 
exist deterministic laws, but not for other variables. This consideration 
leads, in the extreme case, to the conventionalistic assertion that 
through the introduction of certain variables, the law of causality can 
be made valid; this assertion, however, states nothing at all about 
nature but is merely a "de fi nition of the state of aflairs.” That is to say, 
in the language of Cassirer, that through the introduction of an ap- 
propriate concept of the "object” one can always bring about the 
validity of a law of causality. Like the positivists, Cassirer escapes this 
conventionalism only by requiring that the causal laws be “simple.” 
The concept of simplicity remains exactly as vague in his case as in the 
case of the positivists. And the requirement of the "existence of natural 
laws,” which, according to Cassirer, constitutes the real content of the 
general law of causality, involves that indefinite concept of “simplicity” 
in a very fundamental way. The assertion that the general law of 
causality can be expressed only vaguely was one of the essential theses 
of my book “The Law of Causality and its Limits,” which was there- 
fore termed “hyperskeptical” and “antiphilosophic” by many. Exactly 
this character, however, is possessed by Cassirer’s conception of 
causality, which I have therefore designated approvingly as “dis- 
integrating.” Since Cassirer understands by determinism only the re- 
quirement that simple laws exist in nature, he cannot find any con- 
tradiction, of course, between this requirement and modern quantum 
mechanics. In place of the laws of classical physics we have there other 
equally exact laws. Cassirer’s presentation is therefore not a criticism 
of modem physics from the standpoint of “philosophic determinism,” 
but neither is it an attempt to improve "philosophic determinism” with 
the help of modern physics. What Cassirer does is to conduct an in- 
vestigation on the question of how the mles and laws of physics have 
changed their form in recent years because of quantum mechanics. 
This investigation is carried out with a thorough knowledge of the 
subject. Except for a few obscure points, it is in agreement with mod- 
em physics and, naturally, can never come into contradiction with its 
principles. For Cassirer does not set up any philosophic deterministic 


176 



do^erminism and indeterminism in modern physics 

laws as judges over the principles of physics; rather, he is prepared to 
consider every exact law as fulfilling the deterministic postulate. 

Surely this conception is completely tenable, and surely it is more 
useful for understanding modem physics than are the attempts of many 
philosophers to formulate the law of causality more precisely and then 
to interpret physical theories so that they fit into this scheme. An ex- 
ample of such a method is the well-known, in many respects sagacious, 
book by the physicist and philosopher Crete Hermann, “The Philo- 
sophic Basis of Quantum Mechanics.”* She starts out with Kant’s 
special formulation of the law of causality and tries to formulate the 
principles of quantum physics in accordance with it, in such a way that 
for every process there exists another one which the first follows ac- 
cording to some mle. She is then forced, however, to introduce as 
“cause” a process which, aside from the effect for the sake of which it 
was introduced, cannot be observed, so that the law of causality be- 
comes a mere tautology. Such a fraitless procedure is avoided by 
Cassirer through his very general comprehension of determinism. 
Nevertheless, many distinctions between classical and modem physics 
are perhaps lost by it. To express the law of causality, following 
Laplace, as the possibility of predicting future processes is also fmitless 
unless one says how the states of a physical system are to be described. 
One can try, however, to formulate the possibility of prediction in ac- 
cordance with experimental possibilities, on the understanding that 
one describes only such processes as can be actually carried through. 
For example, one can say that, according to classical physics, by a suf- 
ficient refinement of the aiming mechanism one can strike the center of 
a target to within any desired degree of approximation; whereas, ac- 
cording to the quantum theory, if one bombards the target with elec- 
trons it wall never be possible to suppress their scattering below a cer- 
tain degree. We can therefore say in a certain sense that in the field of 
observable processes the new atomic physics is no longer deterministic 
in the same way as classical mechanics. 

In its general features, Cassirer characterizes quite appropriately 

*G. Hermann, Die naturphttosophUchen Crundlagen der Quantenmechanik 
(Berlin, 1935). 


177 



modern science and its philosophy 

the change in physical laws due to modem physics by saying that now 
it is not die concept of “thing” but the concept of “law” that stands in 
the foreground, and that the so-called physical reality, the physical 
object, is only created by the law which we obtain from observation. 
He says: 

We no longer deal with a being, self-contained and absolutely de- 
termined, from which we directly read ofiE the laws, and to which we can 
attach them as its attributes. What really forms the content of our empirical 
knowledge is rather the aggregate of observations which we group together 
in a certain order, and which we can represent through theoretical concepts 
of laws according to this order. 

As far as the dominion of these concepts extends, so far extends our 
objective knowledge. There is “objectivity” or “objective reality” because, 
and in so far as, there are laws— not conversely. From that it follows that we 
cannot speak of a physical "being” otherwise than subject to the conditions 
of physical cognition, including its general conditions as well as those special 
conditions that are valid for its observations and measurements.* 

Reading such deliberations, one comes to think that Cassirer ac- 
cepts completely the positivistic conception of the quantum theory, ac- 
cording to which concepts like “position” or “velocity” of a particle can 
be used only under certain experimental conditions, while the formulas 
of physics only give directions for bringing such observations into 
relation with one another. 

The reader of Cassirer’s book is further strengthened in this opinion 
when he comes to the statement; 

If it turns out that certain concepts, like those of position, velocity, the 
mass of a single electron, can no longer be filled for us with a definite 
empirical content, they must be eliminated from the theoretical system of 
physics, no matter how important and fmitful their contributions may have 
been. 

Often Cassirer even employs a terminology that is linked with the 
nineteenth-century positivism of Kerre Duhem: 

We choose the concepts in such a way that through them the phenomena 
will be described as completely and unequivocally as possible, that through 

* E. Cassirer, op. cU., p. 164. 


178 



determinism and indeterminism in modern physics 


them the phenomena will be preserved. This requirement of ri 

tt>aiv6na>a goes back to the dawn of scientific physics. 

However, when Cassirer tries to describe the real role of such 
concepts as particle, position, and velocity in quantum physics, he 
expresses himself rather vaguely. He presents the more or less provisory 
formulations of various physicists— Schrodinger, Heisenberg, Dirac— 
but does not decide in favor of any one of them as logically the most 
satisfactory. It seems to me that it is best to adopt the latest formula- 
tions of Niels Bohr, which he gave at the second Congress for the 
Unity of Science, held in Copenhagen in 1936, and which are fully 
compatible with the formulations of logical empiricism. Bohr said there 
quite clearly that concepts like “position of a particle” and “velocity of 
a particle” are expressions of everyday language which can be used in 
atomic physics only under special experimental conditions; moreover, 
"position” and “velocity” can be used only under conditions which are 
mutually exclusive. The term “particle” never occurs in atomic physics 
equipped with all the properties that it has in everyday language and 
in the physics of macroscopic mechanical processes. In the descrip- 
tion of many experiments there does occur a particle which sometimes 
has a definite position, sometimes again a definite velocity. The fact 
that in such cases the term “particle” is used at all is due to the con- 
nection with the motion of large bodies. Actually, however, this term 
is used in a somewhat different sense, or, more exactly, according to 
different syntactic rules. 

From the lectures that were held at the aforementioned Congress, 
it is clear that in the complementarity conception of atomic physics 
it is only a question of introducing a new syntax for the words “posi- 
tion” and “velocity of a particle” which is different from that of every- 
day language.' In this connection Strauss pointed out that it is not at 
all a question of introducing mysterious new objects like “particles 
without a definite position.”' All these lectures made it plain that, 
although there may exist a difference between the view of Bohr and 

’ See the lecture of M. Strauss. 

'See, besides the lecture of N. Bohr, those of M. Schlick, V, Lenzen, and 
myself. 


179 



modern science and its philosophy 


that of the adherents of logical empiricism with regard to the applica- 
tion of the complementarity principle to biology and psychology, there 
is certainly no difference in their ideas on the meaning of comple- 
mentarity in physics. 

It seems to me that Cassirer started out with a formulation of the 
complementarity conception in atomic mechanics that was somewhat 
vague, both from a physical and from a logical standpoint. This defect 
crops up now and then also in bis general discussion. It can easily be 
shown that there are many places in Cassirer’s book that cannot be 
understood from the standpoint of logical empiricism, a fact which is 
connected, in my opinion, with his leaving the scientific analysis pre- 
maturely and going over to metaphysics. One gets an inkling of this 
transition in statements such as: 

But a concept like that of material point, from the very nature of the 
matter, can never be understood as the copy of a physical object; it is a 
"form” the meaning and content of which consist in their usefulness for 
the theory, in their ability to lead to simple and rigorous laws for phe- 
nomena. 

What does it really mean to call the material point a “form”? Evi- 
dently it means in the language of physics that the statements in which 
the term “material point” occurs must have a definite syntactic form in 
order to be suitable for the representation of observations, and that 
this syntactic form in quantum mechanics is no longer the same as in 
classical mechanics and in everyday language. Cassirer, however, does 
not use the word “form” expressly in the sense of “syntactic form.” As 
he uses it, it is a reminiscence of the Kantian terminology, in which 
space and time are “forms of experience.” Here the word “form” is 
taken neither in the sense of “spatial form” as in everyday language, 
nor in the abstract sense in which, say, one speaks of the “form of a 
mathematical equation,” but in a quite specific sense which really 
occurs only in Kantian philosophy and may lead to serious misunder- 
standings. If one calls the material point a “form,” one is making use 
of a language that does not fit well into the scheme of physical proposi- 
tions. To be correct, one ought to say: “ ‘Material point’ is an expression 


180 



d^erminism and indeterminism in modern physics 

which, combined with other words according to definite syntactic 
rules, is appropriate for the representation of observations.” 

It is easy to see that Cassirer, in other places, really expresses him- 
self as if, behind the world of relations which the theory sets up among 
observations by the help of its symbols, there existed another “real” 
world which we can approximate only imperfectly. 

Immediately after the statement quoted above, that outside of the ^ 
connections among observations one cannot speak of a “physical 
being,” Cassirer says: 

This “being” has thus lost its ultimate permanence. It is to some extent 
included in the process of physical cognition and is to be considered only 
as a limit to which this process tends, but which is never quite reached. 

The “real world,” that characteristic fiction of every school philos- 
ophy, remains in Cassirer’s conception as a “limit.” It would be con- 
sistent, however, to take this role from it also, for even in this respect 
one cannot speak of it in a scientific way. 

No sooner has Cassirer spoken of this “limit” than he again makes 
statements in a quite positivistic sense. One sees here very clearly the 
disintegration of the school philosophy, which, however, has still left 
untouched a certain dark background, namely that "limit” and those 
“forms.” 

That tliis (in my opinion not entirely consistent) critical attitude 
toward metaphysics prevents him from representing the scientific 
sense of quantum physics with complete clarity is to be seen in several 
places in Cassirer’s book. 

Cassirer represents the facts stated by the Heisenberg uncertainty 
principle in the following way: 

According to the conditions under which the observation is made, the 
object shows us to some extent a difierent aspect. We obtain, according to 
the choice of the measuring instrument and the use we make of it, various 
pictures of the event. No single observation can open up and hold out to us 
at one time the totality of possible aspects. Through every particular measur- 
ing arrangement certain features of the event are, so to speak, screened from 
us, as for example the wave nature or the particle nature of light, whereas 
others are brought out in their place. What the thing is in an absolute sense, 


181 



modern science and its philosophy 


outside of the circumstances of observation as realized in the various experi- 
ments, is something about which we no longer obtain an answer. 

Through this mode of expression Cassirer depicts the situation in 
quantum mechanics as though it were a question of things that are 
absolute, but that cannot be comprehended in all their aspects with one 
measuring arrangement. In this way he brings into physics terminology 
from the philosophy of transcendental idealism, which is entirely for- 
eign to Bohr’s complementarity doctrine in its physical form. Quan- 
tum physics says only that with certain experimental arrangements 
concepts like “particle with a definite position” or “particle with a 
definite velocity” can be defined. In other words, the physical processes 
that occur with these experimental arrangements can be predicted 
through statements in which one refers to “a particle with a definite 
position” or “a particle with a definite velocity,” but there is no ar- 
rangement for which one can predict processes through statements 
involving “a particle with a definite position and velocity.” This, how- 
ever, does not mean that there are particles of which, because of the 
defectiveness of our apparatus or because of malicious natural laws, we 
cannot measure all the characteristics (position and velocity); it means 
rather that such combinations of words as “a particle with coordinates 
X, y, z, and velocity components v^, v” must not be introduced into 
the language of physics. If we were to say that the things correspond- 
ing to such combinations of words nevertheless exist as absolute, but 
unknowable, things, we should be going over into pure metaphysics 
and destroying every bond with experience, which is surely not Cas- 
sirer’s purpose. 

Cassirer thus characterizes often very pertinently the scientific- 
logical structure of the laws of quantum mechanics, but then he al- 
ways formulates them again in the language of idealist philosophy, 
thus robbing them of their clearly delineated scientific meaning and 
opening the door to misinterpretations in the direction of an abso- 
lutistic metaphysics. 

That Cassirer basically does not accept this metaphysical interpreta- 
tion of the quantum theory is to be seen from the determination with 
which he rejects the opinion that from this theory any conclusions can 


182 



fl^terminism and indeterminism in modern physics 

be drawn in favor of free will or even of moral responsibility. Much 
more clearly than many physicists have done, Cassirer sees through 
the deceptiveness of all such arguments and characterizes them very 
appropriately. He says, for example: 

In itself it would be very bad for ethics and its dignity if it could not 
maintain authority except by watching for gaps in the scientific elucidation 
of nature and, so to speak, creeping into these gaps. 

In these words Cassirer ably characterizes the repeated attempts of 
philosophers and many physicists to use the lacunae in science for the 
introduction of supernatural factors. 

At another point he says: 

If the idea of ethical freedom were threatened by these ideas [of 
rigorous natural laws], it could get no help from quantum mechanics. As far 
as this problem is concerned it is immaterial whether we think that a natural 
phenomenon is governed by rigorous dynamical laws or whether we assiune 
merely a statistical regularity. For even from the latter standpoint it would 
be determined to such an extent that the ostensible freedom, the free will, 
could find no refuge in it. An act which, from the physical standpoint, is 
to be regarded as not completely impossible, to be sure, but as improbable 
in the highest degree, is one that need not be taken into consideration in 
the domain of our will. 

And quite tersely and precisely: 

The problem of nature and freedom remains the same whether we take 
the general laws forming the concept of nature as dynamic or as statistical 
laws. 

This separation of the question of natural laws from the question of 
ethical freedom is treated by Cassirer in a way that is very similar to 
that of Schlick in his Froblems of Ethics* which was attacked as being 
extremely positivistic. Also the rejection of the use of gaps in the laws 
of physics for introducing spiritual factors arises from positivistic lines 
of thought and is to be found in a quite similar form in my book “The 
Law of Causality and its Limitations.’' 

^M. Schlick, Fragen der Ethik (Vienna; Springer, 1930); tr. by David R 3 nain 
as Problems of Ethics (New York: ]^ntice-Hall, 1939). 


183 



modern science and its philosophy 


/ 

This attitude of Cassirer toward the question of the relations be- 
tween quantum mechanics and ethics must be valued all the more 
highly in view of the fact that there have been many physicists who 
supported enthusiastically this misuse of the quantum theory, and 
sometimes even initiated it. As a contrast to Cassirer’s rigorously sci- 
entific argument on this question I should like to bring up a few 
sentences from a lecture by the famous English physicist J. H. Jeans. 
He says: , 

The plain average man . . . believed, among other things, that he was 
free to choose between the higher and the lower, between good and evil, 
between progress and decadence. To many, Victorian science seemed to 
challenge all such beliefs. It knew nothing of higher nor lower, progress nor 
decadence; it knew only of a vast machine, which ran on automatically and 
of its own inertia, as it had been set to run on the first morning of the crea- 
tion . . . VVe now begin to think that this challenge was a mistaken one, 
that the universe may be more like the untutored man’s common-sense con- 
ception of it than had seemed possible a generation ago, and that hu- 
manity may not have been mistaken in thinking itself free to choose between 
good and evil, to decide its direction of development, and within limits to 
carve out its own future.® 

At the end of his book Cassirer indicates of what use his philos- 
ophy of quantum mechanics may be. He believes that the change of 
viewpoint between wave theory and particle theory which the quan- 
tum theory brings about within physics is analogous to the change 
of viewpoint that takes place when one goes over from scientific con- 
siderations to ethical or esthetic ones. This analogy has been suggested 
also by many physicists, such as P. Jordan and Crete Hermann, and 
even by Bohr, although very cautiously and with many restrictions. 
Whether one looks upon it as deep and fruitful or as merely super- 
ficial is more a question of guessing future developments than of sci- 
entific argument. 

To summarize: Cassirer’s book is to be welcomed from the stand- 
point of logical empiricism as a highly successful attempt to continue 
the adjustment of the traditional idealist philosophy to the progress of 

® J. H. Jeans, “Man and the Universe,” in Scientific Progress (London; G. Allen 
and Unwin, 1036), pp, 37 ff. 


184 



^terminism and indeterminism in modern physics 


science, which in my opinion can end only with the complete disinte- 
gration of the tradiHongl philosop hy. The ingenious arguments of Cas- 
sirer, presented in clear, understandable language, will be read with 
great benefit by every physicist and will be useful in correcting many 
misinterpretations of modem physics. For the adherent of the school 
philosophy the book signifies, h'ke many previous writings of Cassirer, 
a way out of an impasse. 


185 



CHAPTER 



how idealists and materialists view modern physics 

% k ME know today that nature can be described and understood 
\i\MQOt “mechanistically” but only through abstract mathemati - 
f Y oaI foTinuI as. Great significance has been attached to this 
revolution for the philosophic world view. The argument is that since 
a mathematical formula is something purely mental, the world can no 
longer be understood in a materialistic sense. Materialism must be 
superseded by idealism. The phvsi c^ of the twentieth cent ury is a vic- 
tory for the “spiritualistic” or, as it is sometimes less clearly expressed, 
i dealistic wor ld view. 

This viewpoint has its representatives among both the id ealists and 
th ^ material ists. The former rejoice at the unexpected aid they have 
received for their world view from the progress of science itself. The 
latter blame modem physics for abandoning the paths of progress 
marked out by G alileo and N ewt on and promoting the return to the 
dark views of the Middle Ages. 

We will cite from the writings of a few English and German authors 
to illustrate the point of view of the idealists, and from the works of 
Soviet authors, the apprehensions of the materialists. 

In his well-known book The Mysterious Universe, the famous Brit- 
ish physicist. S ir James Jeans say s, for example: 

The signal for the revolution was a short paper which Einstein published 
in June 1905. And with its publication, the study of the inner working of 

186 ^ 



4)ow ideolists and materialists view modern physics 

nature passed from the engineer-scientist to the mathematic ian . . . The 
essential fact is simply that all the pictures which science now draws of 
nature, and which alone seem capable of according with observational fact, 
are mathematical pictures . . . Nature seems very conversant with the 
rules of pure mathematics, as our mathematicians have formulated them in 
their studies, out of their own inner consciousness and without drawing to 
any appreciable extent on their experience of the outer world . . . Our 
remote ancestors tried to interpret nature in terms of anthropomorphic con- 
cepts of their own creation and failed. The efforts of our nearer ancestors to 
interpret natme on engineering lines proved equally inadequate ... It 
would now seem to be beyond dispute that in some way nature is more 
closely allied to the concepts of pure mathematics than to those of biology or 
of engineering ... In any event, it can hardly be disputed that nature and 
our conscious mathematical minds work according to the same laws. She 
does not model her behavior, so to speak, on that forced on us by our whims 
and passions, or on that of our muscles and joints, hut on that of our thinking 
minds , . , The concepts which now prove to be fundamental to our 
understanding of nature . . . seem to my mind to be structures of pure 
thought, incapable of realization in any sense which would properly be called 
material . . . To my mind, the laws which nature obeys are less suggestive 
of those which a machine obeys in its motion than of those which a musician 
obeys in writing a fugue, or a poet in composing a sonnet . . . The universe 
can be best pictured ... as consisting of pure thought, the thought of 
what, for want of a wider word, we must describe as a mathematical 
thinker.^ 

The physicist and astronomer Jeans has here made use of this turn 
from a “mechanistic” to a " mathemati cal" understanding of physics to 
favor a religiously tinted metaphysics. When the professional German 
philosophers, on the other hand, deal with this revolution in science, 
we find them erecting a “scientific” metaphy sics on this foundation. 

To illustrate this tendency we will quote a few passages from 
B. Bavink’s Science and God. The author stresses that the fundamental 
laws of physics are t oday statements about probab ility. 

But a TTiathfliq^ tical probabil ity is not a physical reality like a tempera- 
ture or field strength or what not. With this new interpretation, the whole 
material .notion nf suhstanca disa ppears in our h ands. What remains then of 

^ J. H. Jeans, The Mysterious Universe (New York: Macmillan, 1930), pp. 106, 
135, 138, 143, 145 ff. 


187 



modern science and its philosophy 


the plain, real, hard, sharp, heavy, etc. matter? A certain probability depend- 
ing on formal mathematical laws, that energy or impulse are observable at 
a certain world-point. 

The physicist of today has learned— an enormous advance from the point 
of view of his world-view— that his atoms or electrons or what not, are no 
longer to be regarded as rigid lumps of reality, from which no path can be 
found into the mental and spiritual sphere; he sees, on the contrary, that 
all these structures are forms in perpetual flux, which are only of interest even 
to him as regards their form. With this view, every variety of materialism is 
superseded.® 

In this mathematical conception of nature, Bavink, like Jeans, sees 
the foundations of a n idealistic philosoph y of physics and hence of 
science as a whole. The only difference is that Bavink uses somewhat 
more technical philosophic terms. 

One task Bavink proposes for an idealistic philosophy of nature is 
the following: 

It still remains to be shown how the material world is to be deduced 
from the purely psychical data. It is obvious that spiritualism has hitherto 
always faffed in this respect. It has never succeeded in deducing even the 
properties of a hydrogen atom from data of this kind. What is new in the 
present situation is the fact that such a proposal no longer appears so 
completely absurd as it did even twenty years ago. For that which hitherto 
presented an insuperable obstacle, namely the “rigid lumps of reality” of 
ordinary atomic theory, has been resolved into pure form, and a mathematical 
form is in itself something psychical, it belongs, as Plato already saw, directly 
in the realm of the Logos, which is behind all things . . . 

There it stands; the hard, cold sober world of matter with its atoms, the 
existence of which is today proven beyond a doubt. It is impossible to pass 
it by, and it is time for all idealists Anally to accept this fact and give up their 
fruitless attempts to avoid it. Matter will only be finally subjugated by mind 
when we are really able to understand it as the product of psychical powers. 
Merely to postulate this as a fact, which is all Aat spiritualism has hitherto 
done, is not of the slightest use; matter and its worshipers, the materialists, 
simply laugh us out of court saying: here is a single atom, the simplest of 
all, the hydrogen atom. Show us what you can do! Show us how we are to 
imderstand it as the product of purely psychical potencies— then we will 
believe you. Now it appears as if spiritualism today can actually pass this 

‘ B. Bavink, Science and God, English translation by H. S. Hatfield (London: 
BeU, 1933), pp. 68, 71. 


188 



how idealists and materialists view modern physics 
* 

test. I will not maintain that it has already passed it, but I believe it to be 
undeniable that it is very close to doing so, and has every prospect of 
success.® 

Some professional philosophers look upon modem physics as a 
direct return to the anthropomorphi c, animistic physics of the Mid- 
dle Ages, as it was practiced by Aristotelian scholasticism. Thus Aloys 
Wenzl, Professor at the University of Munich, says in his essay, 
“Metaphysics of Modem Physics”: 

And in this manner the human struggle for insight in the world wiU 
have described a circle, or more correctly a spiral. The examination of nature 
began with an anthropomorphic representation of the material world, in that 
souls were ascribed to material things, and the relations between them were 
viewed as expressions of psychic relationships of love and hate. The tendency 
to reification led farther and farther away from such early representations, 
freed physics more and more from such images and of necessity led ever 
closer to a mathematical examination of nature. For there are only two 
possible ways of making the facts and the laws of experience meaningful. 
They must either be treated according to psychologic laws or associated 
with the ideal forms of mathematics. Modem physics has followed the 
second method to the very limit. But if more is desired, if assertions about 
their meaning are to be made, the mathematically expressed relationships 
must be explained, which would be a return, on a much higher plane, to he 
sure, to the original method, if not in physics, then in metaphysics.* 

When we notice how a philosopher of the twentieth century “ex- 
plains” the physics of his day, we will see that the “idealisti c” explana- 
tion is not so far removed from the “spiritualistic” or mystical. In the 
same essay, Wenzl says: 

It is clear that the concept of matter has changed completely . . . Only 
the m athematical me thod itself actually defines the sphere of the materfal 
world . . . But we can no longer associate an idea of something dead, with 
this material world. If we do wish to make an assertion about its essence, it is 
much sooner a world of elemental spirits which are bound in their relation- 
ships and their formation of wholes to certain rules of the spiritual realm 
which are mathematically comprehensible; or to put it in other words, it is 

’‘Ibid., pp. 93, 95. 

*A. Wenzl, “Metaphysik der Physik von heute,” Wissenschaft und Zeitgeist 
(Leipzig; Meiner, 1935), No. 2, p. 30. 


189 



modern science and Hs philosophy 


/ 

a world of lower spirits whose reciprocal relationships can be expressed in 
mathematical form. We do not know what kind of relationship this mathe- 
matical form signifies, but we do know the form. Only the mathematical 
forms themselves or God could know their inner significance. A very alert 
metaphysician might explain them at best by analogy to known psychic 
relationships.’ 

All these utterances show what great hopes the supporter s of ideal - 
ism have had in the crisis in physics of the twentieth century. But just 
as some political systems adopt idealism as the foundation for their 
philosophy, so the supporters of Marxism acknowledge materialism 
as a foundation for their world view. When we read the writings of 
the spiritual leaders of Russian Marxism, we see that the crisis in 
physics and its utilization as propaganda for idealism are viewed with 
concern and alarm. 

Lenin published in 1908 a book under the title “Materialism and 
Empiriocriticism: Notes on a Reactionary Philosophy.”® In the fifth 
chapter, entitled “The Latest Revolution in Science and in Philosophi- 
cal Idealism,” Lenin speaks of the changes brought about in the con- 
ception of matte r bv the ene r getic and electromagnetic theories. The 
physicists inclined toward idealism often formulate the results of these 
new conceptions as follows: 

The atom is demateiialized. Matter disappears. 

And the Russian philosopher N. Valentinov in his book “Ernst Mach 
and Marxism” (1907) draws the following inference with reference to 
a world view: 

The assertion that a scientific explanation of the world is found in straight 
^latglialism, is now no more than a myth, and a foolish one at that. 

Says Lenin: 

There is not the slightest doubt about the association of the new physics, 
or rather a certain school of the new physics, with the school of Mach and 
other types of a modem idealistic philosophy. 

The nature of the crisis in modem physics consists in the overthrow of 

® Ibid., pp. 28, 29. 

® English translation (New York: International Publishers, 1927). 


190 



%how idealists and materialists view modern physics 

old laws and fundamental principles. Objective reality outside of conscious- 
ness is rejected and materiali.sm is replaced by ideaUsm and agnostici sm. Mat- 
ter has disappeared. This is the way the basic and typical difBculty created by 
the crisis is expressed. 

Even in the year 1922 when Lenin was already at the head of the 
Soviet Government of Russia, he was much concerned about the dan- 
gers for materialism that might and did arise out of the crisis, thereby 
imperiling the foundations of Communism. At that time it was the 
relativity theory that played the most important part in the crisis in 
physics. 

On December 3, 1922 Lenin wrote in the journal Under the Ban- 
ner of Marxism: 

We must keep in mind that as a result of the revolution taking place in 
science today, reactionary philosophic schools and tendencies are likely to 
arise. Therefore, the journal Under the Banner of Marxism must be con- 
cerned about this revolution in modem science; otherwise, fighting material- 
ism would be neither fighting nor materialism. 

If the great majority of middle-class intelligentsia stand behind Einstein, 
who is not taking an active part in the fight against materialism, then this 
holds not only for Einstein, but for most of the great scientists since the 
end of the nineteenth century. 

We find the same apprehensions and the same fight if we examine 
textbooks that have been introduced in Soviet colleges for teaching 
materialism. Thus we find in the textbook Dialectical Materialism : ' 

The attempts made to think of .motio n without matter and of force 
without underlying substance are laying the foundation for idealis m and 
p]firipa|j<Tn At the present time we see the furthering of these same idealistic 
tendencies. As a result of their association with Einstein’s theory of relativity 
many are inclined to imagine motion without matter. 

And again: 

In place of the old unchangeable atoms there has appeared a system of 
movin g electron s. Therefore, say the "Machians,” matter has disappeared. 
But actually, more exact principles are replacing primitive physical laws. 

^ M. Mitin, ed., DUdekticheskU Materializm ( Moscow Philosophical Institute 
of the Communist Academy, 1934), vol. 1, pp. Ill, 55. 

" 191 



modern science ond its philosophy 


Yet the followers of Mach say: There is no objective knowledge . . . The 
latest quantum mechanics strengthens the concept of causali tv. and makes 
corrections in the old concept. The Machians, however, declare that causality 
has disappeared. 

In the Russian materialistic literature there is a great deal of criti- 
cism of the new physical theories, particularly quantum mechanics, on 
account of their mathematical and formalistic n ature. In effect, the line 
of approach is much like that of the German and English idealists— 
such as Bavink and Jeans— except that the materialists are naturally 
critical about the replacement of mechanics by mathema tics as an 
“idealistic” trend. I will cite as evidence an article entitled “Chemistry 
and the Structure of Matter” which appeared in a Russian literary 
journal, Krasnaya Nov. The author, Orlov, says: 

Quantum mechanics is today still tmder the spell of the fetish , mathe- 
matics. This means that the method of quantum mechanics is of a formal 
T nathp.matical nat ure. The mathematical pattern permits the building of a 
bridge that unites empirical facts furnished by spectroscopy with the be- 
havior of electrons, atoms, and molecules. But up to now we have no physical 
explanation for the formulas of quantum mechanics. Instead of that, it is 
often proposed to abandon the search for an interpretation of physical laws 
and to replace physical representations with abstract mathematical symbols. 
It is in this that the fetish, mathematics, consists. 

It is remarkable how often scientists with sympathies for idealism, 
philosophers with scientific inclinations but of a spiritualistic back- 
ground, and advocates of materialism as a tool for achieving poUtical 
goals agree with one another in so many essential points. 

We have seen in Chapters 5 and 6 that the transition from mecha - 
nistic to mathem atical physics was the result of the positivistic co n- 
ception of science and that it had nothing to do with the twentieth- 
century tende ncy toward idealism and metaphysic s. 

But it may well be, and is often maintained, that positivistic physics 
•Contains an element of spiritualis m nr idealisTn Positivistic physics 
consists in the last analysis o f propositions about observations o r per- 
ceptions. But then, so it is often said, it asserts something about the 
psychic. It has completely abandoned materialism and has become a 


192 



\how idealists and materiolists view modern physics 

science of the mind, exactly like psychology. In fact, Ernst Mach built 
up all the sciences out of relationships among sensation s. 

Many, therefore, look upon positivistic physics as a varie ty of sub- 
iective ideali.s m. which declares that science can never assert anything 
about the actual world, but only about subjective sense impressions. 

In the above-mentioned book. Materialism and EmpiriocrUidsm, Lenin 
represented Mach’s doctrine as a direct continuation of Bisho p Berke - 
ley’s philoso phy. He also blamed this doctrine for depreciating the 
actual world and for promoting the view that behind the world of 
sense impressions, about which alone science can assert something, 
lies the real world, which, however, is inaccessible even to science. 
According to this view, Mach seeks to strengthen the belief that the 
real world can only be explored through extrascientific or superscien-*^ 
tific sources of knowledge such as metaphysics and religion. 

This view of Mach’s doctrine, however, is completely contradicted 
by Mach’s actual intentions. To be sure, it must be admitted that his 
terminology may sometimes be the cause of such confusion. Also many 
of his disciples were in effect inclined toward idealism, perhaps because 
of pressure to conform to the prevailing ofiBcial philosophy, according 
to which everything smacking of materialism was strictly taboo. 

For Mach himself, however, the sensations with which he built 
up the entire science were in no contradiction to the actual world that 
seemed inaccessible to reason. For him these sensations, which to avoid 
confusion he very often termed "elements,” were the building material 
he wished to use to create a unified system of science; this could be 
accomplished if one crossed from one field of science to another, say 
from physics to physiology. 

For Mach, there was no way in which the difference between the ^ 
apparent world and the real world could be scientifically formulated. 

Yet the gross mechanistic philos ophy stood in no logical contradiction 
to Mach’s conception. The bodies of daily experience with which 
mechanism constructed its world were also nothing else than a complex 
of sensations like sight and taste. The question whether there really 
is a matter (a thing-in-itself) that is different from the sensation has 
no meaning for science, since no experiment can be performed that 


193 



modern science and its philosophy 


/ 


might possibly settle the question. Every proposition about a gross 
material thing can also be expressed about sensations. Thus the whole 
of mechanistic physics can be translated into the language of Mach. 
A logical contradiction exists only between a metaphysically conceived 
Machism, which is then subjective idealism, and a metaphysically con- 
ceived materialism, which accepts only matter as having existence. But 
this doctrine was invented by idealists to refute materialism, and was, 
in any case, hardly representative of the thought of scientificaUy 
minded materialists. 

The school philosophers (see Chapter 4) did not have to wait for 
the relativity and quantum theories of the twentieth century to mis- 
construe the transition from me chanistic to mathematical the ory. They 
succeeded in doing this with the physics of the nineteenth and earlier 
centuries. On more than one occasion diey made use of the explanation 
of M ach’s doct rine as subjective idealism. It was even simpler than 
that. Wherever they saw something mathematical they tried to ex- 
plain it as an idg ^istic f.le.m p.nt inside materialistic physics. 

Hermann Cohen, for example, the leader of the neo-Kantian school 
of Marburg, says that mjchanics has acquired a more spiri tual char- 
acter when such subtle mathematical concepts as the differential cal- 
culus had to be used to formulate its laws. 

In the critical supplement to the seventh edition of F. A. Lange’s 
“History of Materialism” (voL 1, pp. 504 ff.) Cohen says: 

The road of research leads straightway to idealism. Materialism is being 
destroyed at the roots of physical concepts, and it is mathematics that is 
leading in the emancipation, which promises to be a lasting one . . . Reality 
is the Real in mass, force and energy— the reality of infinitesimally small 
quantities. There are no other means by which we could even denote the 
real of mass and force for Newtonians, or the real of electric charge for 
dynamicists; much less can we explain these realities in another way. There 
is no other means than the differentials. The infinitesimal is not only the 
origin of every quantity but also the origin of being itself, of the real . . . 
At the roots of physical concepts materialism was destroyed and the libera- 
tion was achieved by mathematics. The old Platonic union of physics and 
mathematics has proved its eternal strength. The mathematical ideas . . . 
offer the solution of the fundamental question of philosophy— the question. 
What is science? 




194 



^ how idealists and materialists view modern physics 


But we have seen in Chapter 5 that the i dealistic int erpretation of 
science, when taken seriously, is the first step in the return to the 
animistic anthropomorphic science of the medieval scholastics. 

Positivism, on the other hand, particul arly logical positiv ism, pre- 
vents the crisis in the mechanistic philosophy from spreading to the 
scientific world view as a whole. It shows that the abandonment of 
mechanistic physics does not imply the need for a return to the anthro- 
pomorphic physics of the Middle Ages. And right here something has 
happened that seems rather paradoxical. All those that advocate a re- 
turn to pre-Galilean science, whether it be under the name of “ideal- 
ism,” “holism,” or “organicism,” or even under the name of “race” 
or “nation,” fight with great ardor for the r etention of the mechanist ic 
philosophy in the field of physics, and condemn the positivistic physics 
as an aberration. 

Evidence for this can be found by merely turning the pages of the 
“Journal for the Whole of the Natural Sciences,” which was published 
in Germany under the Nazi government from 1935 on. Its purpose was 
to fight agains t positiv i sm in s cience. Its main line of attack was to en- 
courage the struggle of “German” science against “French rationalism” 
and “English empiricism.” Organismic philosophy is generally regarded 
as specifically “German.” 

A few passages from K. Hillebrand’s programmatic article, "Posi- 
tivism and Nature," may be cited here. They will make clear how an 
overestimation of mechanistic physics served in the fight against posi- 
tivism and in behalf of organismic science. 

Mechanism was a planned world picture constructed upon a principle; 
positivism accepts without choice every experience into the sum of its ex- 
periences. It accepts to be sure mechanistic explanations, as the equivalent 
of sense perceptions, but it denies as a matter of course the significance of 
every explanation, and gladly disavows them, for its aim is only description 
... I therefore ask, is not the principle advantage of positivism over 
mechanism tied up with as great a disadvantage? Is it really necessary or 
even pleasing to exchange a pictural concept of a mechanistic explanation 
for a pure mathematical formula, which transcends all perception? The 
breakdown of mechanism into positivism— very interesting as intellectual 
history— is an event, it seems to me, that is at present almost everywhere en- 


195 



modern science and its philosophy / 

tirely misunderstood. And yet the principles of scientific method will never 
be determined without its clarification. 

And now the author embarks upon an inspired glorification of 
mechanistic physics. He says: 

He who understands the running of a machine, say a clock, based on the 
complete dependence of rigid bodies on spatial conceptions, has genuinely 
satisfying knowledge. It is the mistake of pnsiHvi.sm that it is .iblp.J-n take this 
intention of Democritus as just one among other arbitrary hypotheses. It is 
also unfortunate for the development of positivism that it retains the ob- 
jectivistic bias of mechanism and has surrendered its only virtue, its pure, 
clear picturization. 

As is almost always the case with those favoring an idealistic- 
organismic science, the enthusiasm for the mechanistic physics in its 
most obsolete form is coupled with a strong aversion for the applica- 
tion of non-Euclidean and multidimensional spaces. Hillebrand 
continues: 

Mechanis m is Euclidea n science. Relativism of non-Euclidean spaces, on 
the contrary, is the favorite child o f positivi sm. A four-dimensional space or 
a space with curved radii is just as logical for pure abstract thought as the 
Euclidean; indeed the dissolution of space and time into mere abstract mathe- 
matical formulas seems to be a distinct gain to this other type of human 
being. The overwhelming eternal advantages of Euclidean space as against 
these abstract arts ought not to be forgotten. 

The reason for the glorification of “mechanistic” physics by the ad- 
vocates of organicism is that for their argument they need the applica- 
tion of a kind of physics that is as narrow as possible and therefore 
most unsuited for the more involved events. Says Hillebrand: 

It is evident from what has been said above that according to our way 
of thinking mechanism is far superior to positivistic empiricism— so long as 
it does not attempt to explain living matter. Besides, since Science has 
abdicated in this respect, there is nothing lost; . . . the value therefore of 
exact scientific research is not attacked, in so far as it is restricted to “dead” 
matter nature. The human mind possesses two sufiicient types of knowledge: 
the explanatory and the understanding. "Anschauun^ in the explanatory 
sense is the mechanistic explanation of nature, the representation of bodily 
form in a EucUdean space-time relation going far beyond positivistic sense 


196 



^how idealists and materialists view modern physics 


perception and yet conceivable or pictorial in a narrower sense of sight and 
taste sensation. The understanding type of knowledge is “Anschauun^ in 
a wider sense, perception not as sensation, but rather as a palpable human 
event that need not be "sensible.” 

The German word Anschauung has two meanings, both of which 
are used in this quotation. It means, first, “optical perception” or “pic- 
tural representation”; second, however, it means “mental intuition” or 
“empathic understanding.” This ambiguity makes the word Anschauung 
a favorite term in idealistic metaphysics. It provides a philosophic 
basis for the “intuition” of the totalitarian leader. 

The tendency is very clearly seen to allow mechanistic science to 
pass as the only "explanatory” type of knowledge that is useful for exact 
science, so that it might be easier if necessary to introduce the so-called 
“understanding” type of knowledge (intuition) in the sciences of 
human conduct. 

Logical empiricism, as opposed to this, stresses the unitary nharnrtpr 
of science. It is not interested in splitting human knowledge into 
“mechanistic-explanatoiy” and “imderstanding-intuitive” types. Our 
modem logic of science depicts the factual process of successful knowl- 
edge and scientific rep resentati on. Mechanistic explanation and intui- 
tive understanding both are popular and rather superficial types of 
scientific representation, but by no means particularly profound types 
of knowledge. 


197 



CHAPTER 



logical empiricism and the philosophy of the Soviet Union 

% A #HEN 1 speak of philosophy in the Soviet Union, I mean 
\ /%^only the system that is officially taught in all schools as 
y Y pliilnsophy— ‘‘ Hialpirtical materia lism.” abbreviated to diamat. 
Of course, in the writings of physicists, mathematicians, and biologists 
one can find many remarks associated with the logic of science. These 
are mostly only echoes of the views predominant in European and 
American science. Besides the official diamat, no other consistent con- 
ception of science has developed in the U.S.S.R. If one wishes to dis- 
cuss the features that are characteristic of the intellectual life of Soviet 
Russia, one must speak only about diamat, concerning which there are 
prevalent in European science very unclear and often greatly dis- 
torted views. 

At first sight, it appears that diamat is extremely hostile to the vari- 
ous forms of logical empiricism. This attitude is shown especially by 
the following examples: 

Empiricism is styled in a stereotyped way “crawling Ai^ricism,” 
because it can never rise to the formation of a scientific system. The 
various forms of neopositivism and logical empiricism are all branded 
with the label “ Machis m." and, as such, are sharply condemned. It was 
perhaps an ominous event for the history of philosophy in the U.S.S.R. 
that Lenin set forth his philosophic views in a book directed against 
the Russian followers of Mach and Avenarius— the book Materialism 


198 



logical empiricism and Soviet philosophy 


and Empiriocriticism.^ In 1935 the twenty-fifth anniversary of the ap- 
pearance of this book was celebrated by all philosophical societies and 
journals of the U.S.S.R. Because in it the doctrines of diamat were 
elucidated by being contrasted with the conceptions of Mach, the 
opinion was established in the official philosophy of the U.S.S.R. that 
Machism-was a movement especially hostile to diamat, and hence to 
be attacked vigorously. In reality, Lenin took issue with Machism be- 
cause it is in many respects related to diamat, and he considered it 
especially suitable for him to bring out his own teachings very sharply 
by means of a polemic against it. 

Because in the teachings of Mach everything is built on the per- 
ceptions as elements, Lenin saw in them a degenerate form o f the sub - 
T^^alism nf RfrkHfyi who had denied the re ality of the world 
of experiences and thu s hsd-made room for the acceptance of a super- 
natural world. On the other hand, because of the connection of 
Machism with the enlightenment philosophy of the eighteenth century, 
its predilection for contact with the physical sciences and its aversion 
to the introduction of any anthropomorphic factors or “psychic” tend- 
encies into science, Machism was reproached with having a “mecha- 
nistic narrowne ss” which rendered it particularly incapable of encom- 
passing social and historical events. 

If we ask what is the attitude of diamat toward the movements 
which have arisen from the synthesis of Mach’s positivism and Rus- 
sell’s logic, we need only open the latest textbook of diamat in the 
U.S.S.R. for information.' There we find it asserted, in effect, that the 
newest Machists want to deepen Machism by the use of symbolistic 
methods. They regard science as a game with empty symbols and thus 
make it incapable of embracing the colorful fullness of the real multi- 
form world. Idealism, mechanism , an d logic ism are only three ways of ■, / 
leading people to a fictitious supersensual world and of. r estraining 
them from occupation with the practical questions of the real world. 
These three doctrines therefore, like religion, ar e npinm for the people, 
putting them into a narcotic sleep which shows them a faded picture 

^English translation by David Kvitko (London, 1927). 

^M. Mitin, DUdektlcheskS MaterUdizm (Moscow, 1934). 


199 



modern science and its philosophy 


/ 


of tie real world. Philosophers who teach idealism, mechanism, and 
logicism are in the service of th e bourgeoisie, just li ke the clergy, and 
make their disciples unfit to work for the social reorganization of the 
world. 

While one gets from these statements the impression of a funda- 
mental antipathy, yet scientific and sociologic considerations indicate 
that this attitude of diamat is rather of a polemical and tactical nature, 
and that it must also contain many elements which are closely related 
to the ideas that we represent. 

L ogical empi ricism developed primarily in the struggle against the 
idealistic metaphysics of thf* .-rhonl philos ophy, which with its odd 
mixture of faded theology and ob solete science has fulfilled a very 
definite social function. The main struggle of diam at is also against this 
metaphysics and this function. In a German textbook of diamat we 
find “metaphysics” described as “an examination of the psendn-renl 
surface , without penetrating into the essential.” Since this description is 
in fair agreement with logical empiricism, we must expect to encounter 
still other similarities. 

The main points of a scientific doctrine related to logical empiri- 
cism which we find in diamat are perhaps the following: ( 1 ) Science 
should be “ materialisti c.” but not “mechanistic”; (2) the criterion for 
the truth of a proposition should be only its confirmation in actual life, 
the doctrine of “ concrete tr uth”; (3) the propositions of science are 
to be understood not only from their logical connection with the prop- 
ositions of the previous stages of science, but also from the causal con- 
nection of scientific pursuits with other social processes. The investiga- 
tion of this causal connection is carried on by a special factual science, 
the sociology of science. 

Here we wish to consider only the first two of these points. 

( 1 ) First of all, we must be quite clear as to what diamat means by 
the word “materialism.” What we generally understand by this word 
as used in popular and even in scientific writings is t he conception of 
all natural phenomena, including human evolution. . a.s analogous to a 
machin e. This view would be called by diamat “mechanistic mate- 


200 



' • IkrUtHtomniricism and Soviet philosophy 

rialism” or “mechanism," and is very strongly opposed. If we look up 
the definition of the word “materialism” in the official textbooks of 
diamat, we find, roughly, the following: “By materialism’ is meant the 
conception that science speaks of a world that is completely indepen d- 
<»n t nf any grTiitrarinps c, a world that is neither the creation of a world 
spirit, as the obje ctive idealism of Hegel h olds, nor the creation of the 
individual consciousness, as the subjective idealis m of Berke ley as- 
sumes.” 

From this form of the definition of materialism, which simply es- 
tablishes the objective character of scientific principles, we shall not be 
able to draw any very specific conclusions of the materialistic concep- 
tion. If, however, we observe how this definition is applied in practice, 
we find that all scientific propositions are to contain only terms that 
occur in statements about observable facts. The description of a 
process is of use to science only if all of the observable aspects of the 
process are embraced. In particular, the part played by the so-called 
psychic processes is not to be emphasized one-sidedly; that would lead 
to “idealism.” If, to take an example given by a textbook on diamat, it 
is asserted that the great power station on the Dnieper, the Dniepro- 
stroy, is the product of the engineering plans, the matter is being de- 
scribed idealistically and one-sidedly. The materialist will say: “Besides 
the plans of the engineers, a decisive part is also played by the new 
social organization introduced by the communist revolution, the new 
conditions of the workers, etc.” Everything in the world that is de- 
scribable through intersubjective expressions is called “matter” by 
diamat. 

This is not to say that matter actually has the properties which 
Newtonian mechanics or even the newer physics attributes to mat- 
ter. Such an opinion would be “mechanical materiali sm.” According 
to diamat, every investigation of the world that makes use of inter- 
subjective expressions is an investigation of matter. The properties of 
matter reveal themselves to us only in the course of the development of 
science. They will never be completely known to us as long as there 
are new laws to be discovered. 

This conception comes very close to the viewpoint that science is 


201 



modern science and Its philosophy 


based on an intersubjective language, which Neurath and Camap have 
designated more precisely as the physicalistic language. Just as, for 
physicalism, the biologic or psychologic propositions are “physical in 
the broadest sense," so for diamat the propositions about the develop- 
ment of life and even about human history are propositions about 
matter. However, just as physicalism does not claim that psychology 
can be reduced to actual physics, so diamat does not say that the social 
development of mankind can be reduced to those laws of matter that 
have been discovered by physics. According to diamat, sociology itself 
discloses new laws of matter. 

Diamat seeks, however, to set up quite general laws for matter, 
laws which are to hold for physics as well as for biology and sociology. ^ 
For this purpose it takes the three laws which Hegel formulated for 
the processes of thought, and from which he also made laws for living 
and inanimate nature because he believed that the whole world is the 
product of thought. Marx and Engels turned Hegel’s teachings upside 
down and began by setting up his three dialectical laws of thought 
as the laws for matter. In this way they founded diamat, “dialectical 
materialism.” The three laws are “the unity of opposites, the transi- 
tion from quantity to quality, and the negation of the 'negation.” We 
see that they still wear their idealistic eggshells. Their application to 
reality is often very much forced, and from their consequences there 
results what L. Rougier^ once called “Soviet my sticism." With these 
three laws of dialectics, originating in idealism, diamat often strays 
from the path of establishing the properties of matter through the 
methods of exact research. Today a determined struggle is being 
carried on within diamat against the “trivializat ion" of dialectics. 

Such rather indefinite principles can often serve to order to some 
extent the empirical material in fields that are still only slightly de- 
veloped, like sociology. If, however, they are applied to sciences where 
we possess better ordering principles, they at once reveal their im- 
perfections. 

Because of these dialectical laws, diamat bears within itself the 
germ of jdealis m. Even in the U.S.S.R. it must perpetually struggle 

'L. Rougier, Les Mystiques politiques contempor^es (Paris, 1935). 

202 



logical empiricism and Soviet philosophy 


against "idealistic deviations,” which in recent years have received the 
name "menshevizing idealism” after the political party of the Men- 
sheviks. Diamat wages continually a “war of two fronts” against ideal- 
ism and mechanism without being able to mark out unequivocally the 
boundaries separating it from these two deviations. 

If it carried on this war of two fronts consistently, it would have 
to discard the idealistic eggshell of Hegelianism, the exaggerated opin- 
ion of the significance of the three dialectical laws. On the other hand 
it would have to avoid the desc ription of matter as something exis ting 
objectively— which is also, in the last analysis, an idealistic concep tion 
—and instead would speak of in tersubjective p ropositions. Then it 
would approach more and more closely the concept ion represented by 
lo gical em piricism, especially by the Vienna Circle. For these groups 
carry on the same two-front war, against the i dealistic school phil os- 
ophy and against the belief tha t Newtonian mechanics in its orig inal*^ 
form is a basis of all science. 

Though, therefore, in our opinion, the dialectical laws do not have 
the importance for a modem conception of science ascribed to them by 
diamat, we must nevertheless admit that something in what it calls 
“dialectical thinking” is quite in line with our ideas. 

This “dialectical thinking” is characterized by Lenin in his re- 
marks on Hegel’s works simply as thinking which has the necessary 
elasticity not to stick to a definite scheme, but which builds itself a new 
scheme corresponding to the given stage of development of science. 
This kind of dialectical thinking is demanded also by logical 
empiricism. 

(2) The second point essential for the understanding of diamat is 
the “doctrine of concret e truth." According to this doctrine, the trath 
of a proposition can never be judged by its abstract formulation, but 
only by examining the practical conclusions that can be drawn from 
it. Whether the idealists or the materialists are right can be judged only 
by seeing what consequences the two doctrines have for practical life. 
This conception is related to American pragmatism. 'The textbooks of 
diamat try to distinguish it from pragmatism by saying that pragmatism 



modern science and its philosophy 


always means a “bourgeois,” that is, an individualistic practice, a test in 
the life of the individual— in “business life,” as they often add derisively. 
Diamat understands by test, above all, the test of a principle in social 
life— in revolutionary practice, as they put it. 

From this doctrine of “concrete truth” one can understand the much- 
discussed attitude of diamat toward religion. By religion is never to be 
understood an abstract system of principles of faith. A thing of that 
sort cannot be tested for truth. By “religion” is always meant a con- 
crete institution, as, for instance, the institution of the church. This 
can be investigated to determine whether it has socially desirable in- 
fluences or the opposite. Definitions of religion like “a feeling of one- 
ness with the universe,” “devotion to a higher duty towards humanity,” 
are rejected by diamat. One textbook remarks scornfully that European 
philosophers, on the basis of such definitions, call even communism 
itself a religion. By religion should be understood a concrete organiza- 
tion which seeks to propagate the belief in a supernatural being among 
men and thus to deter them from the struggle against their oppressors. 
From this point of view one must judge the struggle against idealistic 
philosophy an d against Machism and log icism. To this “doctrine of 
concrete truth” Lenin attributed a great importance for the practical 
political struggle. One must never hold fast to abstract formulas such 
as: for the defense of the fatherland or against the defense of the 
fatherland, for parliamentaiianism or against it. One must rather ex- 
amine in every individual case the practical consequences arising from 
such a demand and see whether they are favorable to the goal pur- 
sued— hence, for Lenin, to the rise of the working class to power. 

Lenin, however, applied this doctrine not only to political but also 
to scientific principles. He insisted that propositions such as “Matter 
is infinitely divisible,” or “ Matter is composed of indivisi ble atoms,” 
are never to be labeled as true or false; they are to be judged by their 
practical consequences, which can also change in the course of the 
development of science. 

The doctrine of concrete truth, if it is formulated conceptually, 
and wherever it is applied exactly, is nothing else than the view that 
the truth of a proposition can only be judged if the methods of testing 


204 



logical empiricism and Soviet philosophy 


\ 

it are given. If somebody states a proposition and fails to state the con- 
ditions, observable in practice, under which he would be ready to ac- 
cept it as true, then it is a proposition that is not scientifically applicable 
—it is meaningless for science. With the doctrine of concrete trut h, 
diamat is therefore defending a standpoint which is very closely related 
to that of positivism and p ragmatism. 

The conception held by many representatives of diamat, that 
logistics is only a formalistic game which avoids having to do with 
reality, is perhaps correct in the case of many metaphysically inclined 
logicians. It is certainly not correct for the Vienna Circle, which uses 
logistics only as an ai d to a r a dical empiricism and p ositivism. 

In any case, the doctrine of concrete truth will some day be ap- 
plied in the U.S.S.R. also to tlie teachings of science. Then it will be 
said: In our time it is no longer appropriate to embrace the new empiri- 
cal and positivistic groups with the idealistic school philosoph y in one 
concept, “the bourgeois conception of science.” The patterns that 
Lenin set up for a concrete situation of struggle should not be regarded 
as general patterns, suitable for the representation of scientific develop- 
ment. It will then turn out that there are very fundamental ties between 
diamat and logical empiricism. 

An analysis of the present situation leads to the conclusion that 
to designate the lo gical empiricis m of today, or logistic neopositivism, 
as “idealistic” or “mechanistic” Machism would be the same sort of ab- 
stractly schematic conception as if one were to label diamat “Hegelian 
idealism” because of the historical connection with Hegel, and for that 
reason to reject it. 

The creative scientific work, particularly in chemistry, physics, 
and biology, that enjoys favorable conditions for development in the 
U.S.S.R. and is in a state of rapid growth, still has very little practical 
effect on diamat. In this situation lies the danger for diamat, which 
may develop in isolation from science like the European school philos- 
ophy, which also claims to give direction to science but succeeds only 
in becoming more and more estranged from science, and consequently 
languishes. 

If, in the U.S.S.R., diamat will strive to cooperate with concrete 


205 



modem science and its philosophy 


/ 


science, those tendencies in it that point toward logical empiricism will 
be strengthened. It will be obvious that the two-front war against ideal- 
ism and mechanism can really be carried on consistently only from 
the standpoint of critical positivism; otherwise one will surely slide 
again into metaphysics to the left or to the right. 


206 



CHAPTER 



why do scientists and philosophers so often disagree 
about the merits of a new theory? 

I F, in seeking an answer to the question that heads this chapter, we 
put the preliminary question, “Do they really disagree?” my answer 
is: At the beginning they do, mostly, but by and by the disagree- 
ment weakens and finally the philosophers come to agree too com- 
pletely. Frequently just at this moment the physical theory in ques- 
tion turns out to be doubtful to the physicist. He advances a new the- 
ory and the whole cycle of disagreement and agreement begins again. 
If we succeed in understanding this periodically recurrent cycle we 
have performed a great step toward the understanding of the interac- 
tion between science and philosophy. 

The divergenc es between physicists and philosophers have become 
very clear recently. We have only to glance at the discussions about 
space, time, and causality connected with relativity a nd the quantum 
theory to see this. There are a great many people who believe that 
these divergencies are characteristics of our twentieth-century physics. 
In order to counteract this erroneous impression I shall start by dis- 
cussing the attitude of scientists and philosophers toward the CopCT- 
nican world system at a time when it was news. My point is that the 
dispute of that time was of the same character as the dispute of toda y. 

Copernicus published his system in the middle of the .sixteeiilh 
QgQtuiy. A century afterwards this system was condenmed by the 


207 



modern science and its philosophy 


Ro man Inquisition as “pTiiio snphinally faisfi .” During this century the 
Copemican system was taught in the universities and presented in 
the oflBcial textbooks as a remarkable achievement in science which was 
—unfortunately— “philosophically false.” This attitude is illustrated by 
some sentences from a textbook of astronomy of 1581 written by the 
Jesuit C. Clavius: 

One may doubt whether it would be preferable to follow Ptolemy or 
Copernicus. For both are in agreement with the observed phenomena. But 
Copernicus’ principles contain a great many assertions that are absurd. He 
assumes, for instance, that the earth is moving with a triple motion which 
I cannot understand. For according to the philosophers a simple body like 
the earth can have only a simple motion. 

After setting forth a number of arguments of the same type the author 
concludes: 

Therefore it seems to me that Pto lemy’s geocentric doctr ine must be pre- 
ferred to Copernicus’ doctrine. 

In spite of the agreement with the observed facts the Copemican 
system had to be rejected because it was in contradiction with certain 
principles which were regarded as firmly established. For instance, a 
simple body like the earth can have only a simple motion, such as a 
rectilinear or circular one. Or, to quote a second principle of the same 
type: We see that every piece of earth falls downwards along a straight 
line; therefore the earth as a whole cannot possess a circular motion. 
For it was an established principle that to every particular kind of 
matter there corresponded a particular type of motion. 

All these principles were parts of Aristotelian physics. 'They orig- 
inated from generalizations of observation, just like any physical 
theorems. 'They belonged, however, to an earlier state of physical sci- 
ence. At the time of Copernicus they were already in a state of “fossili- 
zation”; they were believed to be eternal tmths which could be derived 
from pure reason. Every statement of science that was in disagreement 
with these principles of Aristotelian physics was called “philosophically 
false.” In this sense the Copemican system could be declared “mathe- 
matically true” but “philosophically false.” This meant only that it was 



'disagreement between scientists and philosophers 

in agreement with the observed facts but in disagreement with the 
principles of Aristotelian philosophy or physics— physics being a part 
of philosophy. 

It may be objected that this was the opinion of a Jesuit and ortho- 
dox believer in scholastic philosophy. I shall quote, therefore, the state- 
ment of a very progressive man of that time. Fr ancis Bac on has been 
called in the textbooks of philosophy the very father of empirical 
science. He says, in 1622: 

In the system of Copernicus there are found many and great incon- 
veniences; for both the loading of the earth with a triple motion is very 
incommodious and the separation of the sun from the company of the 
planets with which it has so many passions in common is likewise a diffi culty 
and the introduction of so much immobility into nature by representing the 
sun and the stars as immovable ... all these arc the speculations of one, 
who cares not what fictions he introduces into nature, provided his calcula- 
tions answer. 

If we compare this judgment of an empirical philosopher with that 
of the follower of Aristotle we perceive this difference: the self- 
confident statements of Aristotelian physics now have faded into rather 
vague statements of so-called common sense. We no longer derive from 
profound metaphysical principles the conclusion that the sun must 
possess the same type of motion as the planets, since it is of the same 
nature. It is just a “difficulty” to separate the sun from the company 
of the stars, since they look so similar. 

The philosopher’s attitude toward physical science had, however, 
remained essentially unchanged: physical theorems that are in con- 
tradiction with certain established general principles have to be re- 
jected, even if these physical theorems are in agreement with all ob- 
served facts. The scholastic philosopher just as well as the advocate 
of empiricism upheld the distinction between “ scientific truth ” and 
“philosophic truth ." The truth of a physical theorem in the first sense 
has to be checked by experiments. The truth in the second (the philo- 
sophic) sense depends upon whether the theorem is compatible with 
certain established principles. 

In this sense the Copemican system was declared to be “mathe- 


209 



modern science and its philosophy 


matically true” but “philosophically false,” And this severe judgment 
has been passed again and again by philosophers upon new physical 
theories. Let us direct our attention to Bacon ’s characterization of 
Copernicus’ personality: 

Copernicus was a man who did not care what fictions he introduced into 
nature provided liis calculations answer. 

We cannot help remembering how many philosophic reviewers have 
charged the authors of rec ent physical theo ries (particularly relativity 
and quantum theory), with the same thing. We can perhaps under- 
stand this divergence in the attitude of philosophers and physicists 
most clearly if we examine the example of Newton’s phy sics. For in 
this case we are able to pursue the fate of a great theory from its birth 
to its death. 

As a starting point I quote a judgment on Newton by a great 
philosopher who was Newton’s contemporary— Bishop Be rkeley. The 
point he makes is again the difference between two ways of judging 
a physical theory: either by its agreement with general principles (the 
philosophic criterion) or by the agreement of its consequences with 
observations (the scientific criterion). In his book The Analyst, which 
is devoted mostly to a criticism of Newton’s doctrine, Berkeley puts 
die rhetorical question: 

Whether there can be science of the conclusion when there is no evi- 
dence of the principle? And whether a man can have evidence of the prin- 
ciples without understanding them? And therefore whether the mathe- 
maticians of the present age act like men of science in taking so much more 
pains to apply their principles than to understand them? 

This sounds like a twentieth-century philosopher criticizing Ein- 
stein. It may perhaps be objected that Berkeley was not competent to 
pass judgment on a scientist like Newton. However, a man like 
Leibniz , equally competent as scientist and philosopher, considered 
both Newton’s law of inertia an d his law oLgravit ation philosophica lly 
false and even absurd. 

Two traits in these laws seemed to be incompatible with the es- 
tablished principles of philosophy. According to Newton a moving 


210 



disagreement between scientists and philosophers 

body keeps its with re spect to the emp ty space. This was 

regarded as absurd. How could the empty space exert any action? 
Moreover, the law of gravitation assumed that material bodies at- 
tracted each other at any distance and instantaneously. This action at 
a distance was incompatible with Aristotelian philosophy as well as 
with the “mechanisti c” and "geometr ic" philosophy of Democritus or 
Descartes. For a material body could only be set in motion by contact 
with a second body, bv push or pul l. 

Newton himself did not believe that his force of attraction was a 
causal explanation of the motion of planets. He expected always to 
find a derivation of these laws from general principles which were 
connected with a medium exerting an impact upon the planets. He 
compared himself to a man who could explain the operation of a piece 
of clockwork. Such a man can describe the mechanism by which the fall 
of a weight is transformed into the motion of the hands. But if you ask 
such a man how the weight manages to fall he would be at a loss. 
Nonetheless, you are forced to admit that he has given you a better 
understanding of the clockwork than you had before. 

This attitude is defined in Newton’s famous dictum, “Hypotheses 
non jingo— I don’t set up hypotheses.” This word has been frequently 
misinterpreted. For from our present viewpoint Newton’s law of gravi- 
tation is a hypothesis too. Newton disappointed his most ardent fol- 
lowers in the question of the action at a distan ce as .well as in the 
question of the corpuscular theory of light. He was always convinced 
that this theory was not the negation of the undulatory theory but 
would have to incorporate some elements of the latter. In short, New- 
ton was not a faithful Newtonian. 

The great success of Newton’s physics was based upon the wide 
range of observable facts embraced and by the simplicity and ele- 
gance of the mathematical methods employed. It was justified by its 
consequences, or, to speak in the language of the Middle Ages, by its 
mathematical truth. But the "p hilosop hic trut h” of Newton’s principles 
was regarded as very doubtful by his contemporaries. Not only “pure” 
philosophers but scientists also passed the judgments that these prin- 
ciples were obscure or even absurd. 


211 



modern science and its philosophy 


But presently the confirmation of these principles by the increasing 
range of physical facts that could be derived from them changed the 
attitude of the philosophers too. If we examine the general opinion 
toward the end of the eighteenth century we notice a complete revolu- 
tion. The law of inertia and the law of gravitation were no longer re- 
garded as absurd; on the contrary, they were declared more and more 
to be self-evident, derivable fronLpure_reason, the only way in which 
the human mind can understand nature. 

As an example of this changed attitude we can point to Immanuel 
Kant’.ijJ 'Metapbysical Elements of Natural Science,” which was pub- 
lished in 1786. We find in this book all the theorems of Newton’s Mathe- 
matical Principles of Natural Philosophy, but they are transformed, 
so to speak, into a petrified state. Newton had invented bold generaliza- 
tions in order to cover a large range of facts that had formerly defied all 
at ^atinpgj apprnaeVi All of these general Statements, which 
seemed to Newton’s contemporaries so new, so amazing, so absurd, are 
now quoted as self-evident. Kant claimed to have demonstrated that 
the law of inertia can be derived from pure reason; he claimed that 
the recognition of that law is the only assumption under which nature 
is conceivable to human reason. 

One may say that this was merely the opinion of a philosopher who 
was a product of the German inclination toward a foggy metaphysics. 
But when we look at the great advocates of empirical p h iloso phy in 
the nineteenth century we find almost the same opinion. We may 
choose as an example the British champion of .empirical and m echa- 
nis tic philoso phy in the middle of the nineteenth century, Herbert 
Spence r. In his standard work, Synthetic Philosophy, he expresses him- 
self about the law of inertia. He says: 

This law means t hat motion like m auer jg This inde- 

structibility is not inductively inferred, but is a necessity of thought. For 
destructibility catmot be conceived at all ... it is a pseudo-idea. To say 
that something can become nothing would establish a relation between two 
terms of which one (nothing) is absent from consciousness, which is im- 
possible. 

This was written in 1860. 


212 



* disagreement between scientists and philosophers 

But we may leave the philosophers and examine the attitude of 
the scientists of that period. We soon notice that their attitude is 
strongly influenced by the success claimed by the philosophers. The 
scientists wouJd not exactly say that Newton’s principles of mechanics 
could be derived from pure reason, but they would fervently proclaim 
that no physical theory is satisfactory which fails to prove that the 
observed phenomena are derivable from Newtons laws. Without this 
proof no theory could be regarded as a real step toward the under- 
standing of nature. I quote two striking examples. About the middle 
of the nineteenth century, in 1847, Helmholtz published his famous 
paper, “On the Conservation of Energy.” ^ He was a great physicist who 
was also a great physiologist and psychologist. He said: 

The task of physical science is finally to reduce all phenomena of nature 
to forces of attraction and repulsion the intensity of which is dependent 
only upon the mutual distance of material bodies. Only if this problem is 
solved are we sure that nature is conceivable. 

Perhaps still more impressive are the statements of the well-known 
physiologist Du Bois-Reymond. He gave, in 1872, an address “On the 
Limitations of Natural Science.” This speech was widely discussed in 
the last quarter of the nineteenth century. Du Bois-Reymond said: 

The cognition of nature is the reduction of changes in the material world 
to the motions of atoms, acted upon by central forces, independent of time 
... It is a psychological fact of experience that wherever such a reduction 
is successfully carried through our need for causality feels satisfied for the 
time being. 

Is this not an amazing fact in the history of human mind? As New- 
ton set up his theory the introduction of the central forces of attraction 
was regarded as a particularly weak point of this theory. It was accused 
of requiring the introduction of an element that is philosophically 
absurd. But what happened about a hundred years later? It was claimed 
as a “psychologic fact” that just the same thing— the reduction of a 
group of phenomena to the action of central forces— satisfies our need 
for causal understanding. And the derivation of physical theorems from 

^H. von Helmholtz, Vber die Erhaltung der Kraft (Berlin, 1847). 


213 



modern science and its philosophy 


the action of these forces, which were formerly condemned as un- 
conceivable, was now the guarantee that nature is conceivable. 

What is the point of all these considerations? By examining the 
changes in the appreciation of Newton’s laws we are able to find out 
and to understand the origin and the for mation of established phi lo- 
sophic princi ples. Both the law of inertia and the law of gravitation 
originated as physical hypotheses that enabled the physicist to describe 
and predict a large group of observable phenomena in a very con- 
venient way. These laws were justified by the success of this enterprise 
—that is to say, by their effects— but they could not be recognized as 
compatible witli established philosophic principles. They were, if we 
apply the language of the Church in its struggle against Copernicus, 
^“philosophically false” and merely “mathematically true.” 

But what was the situation in the middle of the nineteenth century? 
Now, the same laws, the law ofjR^ti^^weUas the law of gravitation, 
became themselves established philo sophic p rinciples, with which all 
physical theorems had to be in agreement. A physical theorem was now 
by definitionJ^hilosQphicaUyJbrue” if it could be derived from Newton’s 
laws. We understand now very well that these “established philosophic 
'^principles” are nothing else than physical hypotheses in a state of 
petrifaction. 

It may be asked why we should call it ‘^petrifaction.” A physical 
theory can be changed when new facts are discovered that are not 
embraced by this theory in a convenient way. But a philosophic prin- 
ciple which is derived from pure reason can rigyer be changed or even 
modified. I£ Kant and Sp5ncer_are right, that the principle of inertia 
can be demonstrated by purely mental operations, no future discovery 
of new physical phenomena can bring about any modification of this 
principle. The transformation of a physical hypothesis into a philo- 
sophic principle is therefore a petrifaction of that hypothesis. 

And now it seems very plausible that the philosophic principles of 
earlier periods are of the same origin, ^ristode’s principles of physics 
were originally also generalizations that covered in a convenient way 
a certain group of observed facts. When Copernicus and Galileo ad- 
vanced their new physical theories they were declared to be “philo- 


214 



disagreement between scientists and philosophers 


sophically false.” This meant only that they were in contradiction with 
the petrifactions of Aristotelian physics. 

In the same way we can now understand the widespread claim that 
the theory of relativity and the quantum theory are valuable des crin- 
tions of observed facts but give us neit her a causal u nderstanding nor 
a description of phy sical reality. To put it briefly, they are taxed with 
being only mathematically true.4)ut-philosDphically_ialse or even 
absurd. This means in this case only that they are in contradiction with 
the petrifactions of Newton 's_ physi cs. Or in other words, in twentieth- 
century physics we are confronted with new experimental facts and 
have to change the hypotheses of Newton’s physics. This is possible 
as long as these hypotheses are not petrified. But once Newton’s laws 
are regarded as phil osop hic principles.which.can b e deduced from pur e 
reason they can n o longe r -be- changed. Now every modification of 
Newton’s laws will be "philosophically false.” 

But knowing the origin of philosophic principles we need not be 
terrorized by the verdict "philosophically false.” It means only that the 
new physical laws are in contradiction.with ihLe. old ph ysical law s which 
appear now disguised as philosophic principles with pretensions of 
eternal validity. The old physical theory was a good description of a 
restricted group of facts. But to cover the new facts the old theory 
became inconvenient. It is natural to drop it, if an obsolete physical 
theory does not pretend to be an “eternal philosophy." 

This very simple state of affairs has often been described by the 
pretentious term “crisis of physics,” or even "cris is of scienc e.” 

And now we can answer with a few words the question put in the 
title of this chapter. Why do philosophers and scientists so often dis- 
agree about the merit a ha new theory ? They mostly disagree because 
the new theory seems to be in contradiction to established philosophic 
principles. Moreover— and this is my chief point— this disagreement 
arises from necessity, for the established philosophic principles are 
mostly petrifactions of physical theories that are no longer appropriate 
to embrace the facts of our actual physical experience. 


.2J5 



CHAPTER 



the philosophic meaning of the Copernican revolution 

I N 1543, four hundred years ago, Copernicus died. This year was in a 
certain sense also the year of the birth of the Copernican system. 
His great book, “The Revolutions of the Celestian Bodies,”^ was 
published in the same year. When this book was published, nearly 
fifty years had passed since Columbus discovered America. This event 
was one of Copernicus’ starting points. He refers in one of the first 
pages of his book to this discovery as the final proof of the spherical 
shape of the earth. (Incidentally, he does not mention Columbus and 
says, as a matter of course, that “America is named after her dis- 
coverer.”) The earth was now definitely a globe, and, to speak in terms 
of today, “global thinking” could begin. 

To understand a phenomenon means to interpret our present ex- 
perience as the repetition of a similar phenomenon of the past. This is 
true in science, but it is true as well in history. Today, we understand 
the French Revolution better than our parents did because we are 
contemporaries of the Russian revolution, and we understand the 
Copernican r&voluHnn better than nineteenth-century scientists did 
because we are contemporaries of the Einsteinian revo lution. 

Let us formulate the central problem of all philosophy of science in 
the simplest possible terms. We have to face two worlds: on one hand, 
that of our sensg_£)hseml3eaG, sudi as, in astronomy, the observed 
^ De BevoluHonlbus OrbUm Celestium. 



philosophic meaning of the Copernican revolution 


positions of the stars in the sky; on the other hand, that of the general 
prinHpT^F such as the law of gravitation and the principle 

of relativity. To what extent are these general principles justified by 
those sense observations? 


This sounds simple enough, but to appreciate the immense gap be- 
tween these two worlds means to start grasping the central problem of 
all philosophy of science. Unfortunately, the pedagogic effort of science 
teachers has been often directed towards camouflaging this gap. If, 
however, the very goal of science teaching is to help the students in 
the understanding of nature, the actual depth and width of that gap 
^must be emphasized over and over again. The state of mind acquired 
by the average student of science as the result of inadequate training 
has been largely responsible for the failure to appreciate exactly the 
philnsapbie-meaning of the Copernican revolution. 

Only recently, under the violent impact of twentieth-century 
physics, particularly the theory of relativity, have the eyes of science 
students been opened and has the meaning of the Copernican revolu- 
tion become clear. 


If we look into a typical textbook or listen to an average teacher, 
we learn that before Copernicus, men believed in the testimony of their 
senses, which told them that our earth is at rest, that the planet Jupiter 
traverses in twelve years a closed orbit on the celestial sphere and that 
this curve contains twelve loops. Finally, Copernicus recognized the 
fallacy of this testimony and proved that “in reality" our earth is in 
motion and that the planet Jupiter traverses “in reality” a smooth circle 
around the sun as center. Copernicus exposed the illusions of our 
^nses. Human reason thus scored a clear victory in its struggle against 


This description of Copernicus’ achievement seems to me, con- 
servatively speaking, inadequate. The loops traced by the planets are 
by no means a sort of optical illusion and neither is the immobility of 
the earth. As a matter of fact, the planet Jupiter actually traverses every 
year a loop with respect to a system of reference that participates in 
' the annual revolution of the earth. But the same Jupiter traverses just 
as truly every twelve years a smooth circle with respect to the system of 


217 



modern science and its philosophy 


fixed stars. Neither the loops nor the smooth circle are results of our 
naive sense experience. They are two different diagrams representing 
one and the same set of sense observations. Therefore, the interpreta- 
tion of Copernicus’ achievement as a victory of abstract reason over 
naive sense experience is hardly justifiable. 

However, we meet occasionally a second interpretation which say? 
almost the opposite of the first one. The hard facts of our sense ex- 
perience became more and more incompatible with medieval phi- 
losophy, which had its roots in speculative reasoning rather than in 
sense observation. Copernicus finally decided to overtlirow the obsolete 
doctrines of Aristotle and Ptolemy and scored a victory for experience 
in its struggle against pure speculation. As a matter of fact, Copernicus 
was not particularly “tough minded,” if we may use the famous phrase 
of William James to describe the empirical scientist as distinct from 
the "tender-minded” believer in pure reasoning. No new facts had been 
discovered by Copernicus, which had forced him to abandon the 
geocentric doctrine. The astronomical tables calculated from the 
Copernican system were in no better agreement with the observed 
positions of the stars than the previous tables. 

Therefore we have to start from the fact that the Copernican revolu- 
tion meant neither a victory of reason over the illusion of our senses 
nor a victory of hard facts over pure speculation. To be sure, Coper- 
nicus invented a new pattern of description for our observations. His 
genius manifests itself in the beautiful simplicity of this pattern: he 
replaced loops by concentric circles. 

Copernicus died in 1543. The Roman Holy Office did not utter an 
oflBcial judgment on the Copernican system until 1616, seventy-three 
years after Copernicus’ death. This Roman verdict will give us the 
best hint about the philosophic meaning of the Copernican revolution. 
For the verdict considered specifically the philosophic merit of the 
new system. The Copernican theory was called “pliilnsophicaily foolish 
and absurd .” 

But not even Copernicus’ greatest opponents ever doubted that his 
system meant a great advance in astronomy. The general opinion in 
these quarters was that the heliocentric system is “astronomically true,” 



philosophic meaning of the Copernicon revolution 

or as it was sometimes phrased, “mathematically true,” but in any case 
“philosophically false” or even “absurd.” 

We have to do here with a conflict between two conceptions of 
truth. This conflict has existed through the ages and has created quite 
often a great confusion of mind. This double meaning of truth has never 
been dramatized so clearly as by the Copernican revolution and its 
repercussions. To understand and to evaluate this conflict is the great 
lesson we can learn from the history of the Copernican ideas. 

The medieval philosopher St. Thomas Aquinas described very 
distinctly two different criteria of truth: 

There are two ways to prove the truth of an assertion. The first way con- 
sists in proving the truth of a principle from which this assertion follows 
logically. In this way, one proves in physics the uniformity of the motion of 
the celestial spheres. The second way consists not in proving a principle 
from which our assertion can be derived but in assuming our assertion 
tentatively and in deriving results from it which can be compared with our 
^Observations. In this way one derives, in astrology, the consequences of the 
hypothesis of eccentrics or epicycles concerning the motion of celestial 
bodies. However, we cannot conclude in this way that the same assumptions 
cannot be derived, perhaps, from a different hypothesis also.* 

If a statement of astronomy met only the second criterion, the 
agreement with observed facts, it was termed “mathematically true.” 
Only if it met also the first criterion, that is, if it could be derived from 
an evident principle, was it recognized to be “philosophically true.” 
Since Aristotle’s physics was supposed to be derived from evident 
principles, to be philosophically true meant practically to be in agree- 
ment with Aristotelian physics. 

As Copernicus had been anxious lest his system might not be 
philosophically true in this, sense, he feared some hostility on the part 
of theologians who were strict believers i n Aristotelian phi losophy. He 
looked for advice on how to behave in this situation, and strange as it 
may seem to us, the Catholic churchman Copernicus asked a Lutheran 
theologian from Nuremberg how to avoid trouble. The Nuremberg 
scholar, Osiander, answered him in a letter of 1541: 


* Summa Theologlae, 


219 



modern science and its philosophy 


As for my part, I have always felt about hypotheses that they are not 
articles of faith but bases of calculation. Even if they are false, it does not 
matter much provided that they describe the observed phenomena cor- 
rectly ... It would, therefore, be an excellent thing for you to play up a 
little this point in your preface. For you would appease in this way Aristo- 
telians and theologians, the opposition of whom you fear. 

This advice meant precisely tliat Copernicus should not claim 
“philosophic truth” for his system but should be satisfied with a claim 
for “mathematical truth.” 

But Copernicus did not like this compromise. He claimed his system 
to be at least as philosophically true as the Ptolemaic system, and per- 
haps even more so. In this way a conflict flared up, the issue of which 
was a very subtle distinction. Was the Copemican doctrine a true 
description of the universe or was it merely an hypothesis which served 
for calculating the positions of the stars? And how did Copernicus 
himself look upon this question? Most of the scientists of today are 
accustomed to regard every theory as a working hypothesis only, and 
would hardly be prepared to give serious thought to that subtle dis- 
tinction which is rather an issue of the philosophy of science. But if we 
go a little deeper into the logical structure of science, we have to 
recognize that, as a matter of fact, every scientific theory, of whatever 
period, had to meet the two requirements of a “true theory” which were 
already familiar to Thomas Aquinas. In reality, no theory was accepted 
merely because it was a good working hypothesis. In every period of 
the history of science a theory had to be in agreement with the general 
principles of physics. The physicists of the nineteenth century would 
hardly have admitted a theory that was in disagreement with the 
principle of conservation of energy. 

For that reason, practically every theory has been a compromise 
between these two requirements. This is particularly true of the 
Ptolemaic system. We read and hear frequently that the Ptolemaic 
system was in agreement with the Aristotelian philosophy and physics. 
But Copernicus, we are told, disturbed this harmony and advanced a 
theory that would contradict explicitly the laws of medieval physics. 
This was certainly not the opinion of the medieval philosophers them- 


220 



philosophic meaning of the Copernicon revolution 

selves. One of the basic principles of medieval physics was the law that 
terrestrial bodies move in rectilinear paths toward or away from the 
earth while celestial bodies move in circular orbits with the earth 
as center, but the Ptolemaic system assumes that sun and planets 
traverse eccentric circles or epicycles the center of which is not the 
earth. Therefore, the Ptolemaic system could not be regarded as philo- 
sophically true, but at most as a hypothesis that might serve as a basis 
of calculation. 

Thomas Aquinas judged the Ptolemaic system as follows: 

The assumptions made by the astronomers are not necessarily true. 
Although these hypotheses seem to be in agreement with the observed phe- 
nomena we must not claim that they are true. Perhaps one could explain the 
observed motion of the celestial bodies in a different way which has not been 
discovered up to this time. 

The twelfth-century Arabian philosopher Averroes and his school 
emphasized very strictly the philosophical criterion of truth and de- 
clined to ascribe any truth value to the Ptolemaic system. Says Averroes: 

The astronomers start from the assumption that these [eccentric or 
epicyclic] orbits exist. From this assumption they derive results that are in 
agreement with our sense observations. But they have not proved by any 
means that the presuppositions from which they started are, in turn, neces- 
sary causes of these observations. In this case, only the observed results are 
known but the principles themselves are unknown, for the principles cannot 
be logically derived from the results. Therefore new research work is neces- 
sary in order to find the “true” astronomy, which can be derived from the 
true principles of physics. As a matter of fact, today there is no astronomy 
at all, and what we' call astronomy is in agreement with our computations 
but not with the physical reality. 

The common opinion among philosophers was rather that the 
true picture nf the universe cann ot he discovered by the astronomer, 
who is restricted to finding out what hypotheses are in agreement with 
observed facts. If different hypotheses meet this requirement, science 
cannot decide which is true and, as the Jewish medieval philosopher 
Moses Maimonides puts it: 


221 



modern science and its philosophy 


Man knows only these poor mathematical theories about the heavens, 
and only God knows the real motions of the heavens and their causes. 

It is certain, therefore, that before the Copemican revolution no 
theory of the motions of the celestial bodies existed that would meet 
both criteria of truth. There was in every theory a discrepancy between 
mathematical and philosophic truth. Against this background we have 
to interpret the famous dedication letter which Copernicus published 
as a preface to his great book and in which he recommended his work 
to the good will of Pope Paul III. 

Copernicus affirms that he did not advance his new theory of the 
motions of the heavens in a spirit of opposition against the established 
doctrine. His only motive was his conviction that there was no es- 
tablished doctrine. The hypothesis of a circular motion of planets 
around the earth as center did not account for the observed facts, and 
the hypotheses of eccentrics or epicycles were not in agreement with 
the general principles of physics which required uniform circular mo- 
tions around the earth as center. Since no doctrine existed which could 
be regarded as “true” from the philosophic as well as from the mathe- 
matical angle, Copernicus felt free to suggest a new hypothesis assum- 
ing the mobility of the earth. 

This hypothesis accounted for the observed motions nearly as well 
as the Ptolemaic theory of epicycles, but removed some of the epicycles. 
The motions of the planets became now circular orbits around the sun 
as center, except for the epicycles which were necessary to account for 
the inequalities in the motion of planets. In any case there were fewer 
epicycles and more homocentric orbits in the Copemican, than in the 
Ptolemaic, system. Therefore Copernicus claimed that his theory was in 
some sense nearer to the requirements of Aristotelian physics than was 
the geocentric system. The Ptolemaic system was a compromise be- 
tween the two criteria of truth. Copernicus claimed that his system was 
in the same sense a compromise and, as he believed, a better one. 

In any case, Copernicus claimed to give in his theory a true picture 
of the universe, true in every sense of the word. By a strange coinci- 
dence. Copernicus’ book was edited by the same Osiander of Nurem- 
berg whose advice Copernicus did not like to follow. We understand 

«. C 


222 



philosophic meaning of the Copernicon revolution 


now the famous words of the editor s preface, which had been originally 
ascribed to the author himself but which reflect only the editor’s 
opinion; 

The hypotheses of this book are not necessarily true or even probable. 
Only one thing matters. They must lead by computation to results that are 
in agreement with the observed phenomena. 

While Copernicus tried to achieve the compromise by arguing that 
his theory is to a large extent in agreement with the principles of 
Aristotelian phy sics. Galileo Galilei, in his famous Dialogue on the 
Copernican and Ptolemaic Systems of the World, went a good deal 
further in the overthrow of medieval science. He no longer attempted 
to reach the compromise by adjusting his working hypotheses to the 
requirements of the established principles of physics. On the contrary, 
he ventured to adjust the principles of physics to the best suitable work- 
ing hypotheses. This meant dropping the bulk o f Aristotelian phy sics 
and starting a movement in science that led in time to the philosophy 
of science which we would call toda y positivism or pragma tism. The 
two criteria of truth which were for medieval thinkers like St. Thomas 
Aquinas two distinct requirements, have fused more and more into 
one single requirement: to derive the best description of the observed 
phenomena from the simplest possible principles, while these principles 
are justified solely by the fact that they permit this derivation. 

Galileo’s ideas were not brought into a coherent system of proposi- 
tions until Tsaao Nswtn n advanced his celebrated laws of motion in his 
Mathematical Principles of Natural Philosophy. This book appeared in 
1687, approximately 150 years after the popemican rfivnlujjnrL. From 
the Newtonian principles the Gopemican doctrines could be logically 
derived. Therefore, to the believer in these principles, the Gopemican 
system was now true in the full sense of the word, philosophically and 
mathematically true. 

Let us now ask. What did the Gopemican hypothesis look like when 
it was derived from the Newtonian principles? It said that the earth 
is rotating with respect to absolute space and that the planet Jupiter 
traverses smooth circular orbits with respect to absolute space. But 

223 



modern science and Hs philosophy 


Newton himself was very well aware that “motion relative to absolute 
space” has, to use P. W. Bridgman's term, no operational meaning, 
that is, that by no physical experiment can the speed of a body in 
recb'linear motion with respect to al»olute space be measured. 

Therefore, the Newtonian system of principles is not a logically 
coherent system within the domain of physics. Newton himself restored 
logical coherence by enlarging his system of physical statements by 
the addition of some theologic propositions. As we read in Burtt’s book 
on the Metaphysical Foundations of Modem Physical Science; 

CertaiiJy, at least God must know whether any given motion is absolute 
or relative. The divine consciousness furnishes the ultimate center of refer- 
ence for absolute motion. Moreover the animism in Newton’s conception of 
force plays a part in the premises of the position. Cod is the ultimate origina- 
tor of motion. Thus real or absolute motion in the last analysis is the ex- 
penditure of divine energy. Whenever the divine intelligence is cognizant 
of such an expenditure the motion so added to the system of the world must 
be absolute. 

Under the influence of the spirit of the eighteenth century the 
mixing of theology into science began to be regarded as illegitimate. 
Strange as it may seem, by the abandonment of theologic argument the 
Newtonian physics lost logical coherence. Burtt says very correctly: 

When in the twentieth century Newton’s conception of the world was 
gradually shorn of its religious relations, the ultimate justification for ab- 
solute space and time as he had portrayed them disappeared and the entities 
were left empty. 

Therefore the new principles of physics from which the Copemican 
theory could be derived were far from being satisfactory. The “philo- 
sophic truth” of the Copemican system was still a doubtful thing. 

Toward the end of the nineteenth century, F.mst Mach exposed 
very specifically the logical incoherency of tha Nfiwtonign 
as a purely physical system. He claimed on good grounds that the 
principles from which the Copemican syste m was derived are es- 
sentially thenlngic nr meta physical princip les. Mach claimed in the 
nineteenth century, as Averroes had done in die period of the Ptole - 


224 



philosophic meaning of the Copemican revolution 


nasaiLsystem, that we have no true astronomy, if “true” means “derived 
from a coherent system of principles of physics.” ^ 

Mach asked for the remo val of the concept of absolute space- from 
physics and for a new physics which contains only terms which have 
within physics, to speak again with Bridgman, operatiopal 

This program, however, was not carried out until Einstein created 
his general theory of relativity and gravitation between 1911 and 1915. 
This theory, as a matter of fact, was the first system o f-purely physica l 
p rinciples from whirh the Cn pemioan system of planetary motions 
could be derived. But the description of these motions looked now very 
diEerent from the way it had looked as derived from Newton’s prin- 
ciples. The concept of absolute space was no longer present. There- 
fore the statement of the rotation of the earth and of the smooth circular 
orbits of the planets had now to be formulated quite differently. 

From Einstein’s principles one could derive the description of the 
motions of celestial bodies rel ative tn any -system of reference. One 
could demonstrate that the description of the motion of planets becomes 
particularly simple if one uses the system of fixed stars as a system of 
reference, but there was still no objection to using the earth as system 
of reference. In this case, one obtains a description in which the earth 
is at rest and the fixed stars are in a rotational motion. What appears 
to be in the Copemican heliocentric system the centrifugal force of the 
rotating earth becomes in the geocentric system a gravitational effect 
of the rotating fixed stars upon the earth, 

The Copemican system became for the first time in its history not 
only mathematically but also philosophically tm e. But at the same 
moment the geocentric system became philosophically tme, also. The 
system of reference had lost all philosophic meaning. For each astro- 
nomical problem, one had to pick the system of reference that rendered 
the simplest description of the motions of the celestial bodies involved. 

The reception of the Einsteinian revolution by the scientists of the 
twentieth century reminds us in many respects of the reception of the 
Copemican revolution by the scientists of the sixteenth century. This 
comparison might help us to understand the philosophical meaning 
of both. 


225 



modern science and its philosophy 


We may take as an example the way in which Einstein deals with 
the contraction of moving bodies in the direction of their motion. The 
verdict of quite a few twentieth-century physicists was: the theory of 
relativity permits us to derive the observed phenomena from hypo- 
thetical principles but it does not give a physical explanation of the 
contraction. This was an exact repetition of the Roman verdict against 
the Copemican system. For the meaning was: the theory of relativity 
may be “mathematically true” but it is certainly “philosophically false.” 
Now “philosophically false” meant not to be in agreement with New- 
- ^tons principles of phy sics, while in the sixteenth century the same 
expression meant not to be in agreement with Aristotle’s physics and 
Scholastic philosophy. 

But what are the facts affirmed by the Copemican doctrine which 
are still accepted today as true? Copernicus enthusiastically proclaimed 
the sun as the center of the universe and said: 

In the center of the Universe the sun has its residence. Who could locate 
this lamp in this beautiful temple in a different or better place than in the 
center wherefrom it can illuminate the whole of it simultaneously? 

Even if we restrict the meaning of the word “universe” to our 
galactic system, the Milky W av, this universe is not spherical and the 
sun is not located in the center. It has been known for a long time 
that our galactic system has the shape of a lens. Before the distance of 
very remote stars could be estimated, it was plausible to believe that 
our sun, with our earth as attendant, is located in the center of this lens. 
However, in the twentieth century new methods were developed for 
estimating the great distances of remote stars, in large part by Harlow 
Shapley and his collaborators of the Harvard Observatory. In par- 
ticular, Shapley found that our sun is not located near the center of 
that lens, but approximately 30,000 light years away from it. This 
means that the sun with our planetary system is near the edge of the 
lens. According to Copernicus, we inhabitants of the earth have no 
longer the great satisfaction of being the center of the universe, but we 
have at least the small satisfaction of being the attendant of a master 
who has his residence at this center. But according to Shapley, man 



philosophic meaning of the Copernican revolution 


has lost all reasons for complacency. He is not even the attendant of a 
master who occupies the central stage of the universe. 

Copernicus probably believed that the orbits of celestial bodies can 
be described in the best and simplest way by taking the sun as a body 
of reference. In our twentieth century, we know that this caimot be 
true universally. According to Einstein s theory of gravitation, there is 
no "all-purpose system of refere nce.” Copernicus’ suggestion of using 
the sun is practical only if we restrict ourselves to the motions in our 
planetary system. For eve ry particular purpose a particular svsten i' 
may be the most suitable. 

Copernicus did not discover any new fact that could be regarded as 
established for all eternity. But he denied to the earth its former role 
as the only legitimate body of reference, he demonstrated that the sun 
is the most suitable system for a particular purpose, and he cleared 
the way for the great new truth that we have complere fr^pdrYm in,f> iir 
choice of a system of reference. 

The Copernican revolution did not end by replacing the earth as 
master of the Universe by the sun or by absolute space, but it was only 
the first step in a series of revolutions that culminated, so far as we 
know today, in depictin g a de"i »<^»-aHr» nrriar nf the universe in which 
all celestial bodies play an equal part. 


227 



CHAPTER 



the place of the philosophy of science in the curriculum 
of the physics student 

I F we wish to exercise sound judgment about the success or failure of 
science teaching in the general scheme of our educational system, 
it may be a good idea to arrange a hearing. We may then listen to 
people who were compelled to submit to this teaching and later 
achieved such a great reputation in om cultural life that we cannot 
ignore their opinions about the contribution of science teaching to 
general education. 

Ralph Waldo Emerson, who complained that science had become 
abstract and remote from human life, says in his essay on Beauty in 
“The Conduct of Life”: 

The formulas of science are like the papers in your pocketbook, of no 
value to any but the owner . . . There’s a revenge for this inhumanity. 
What manner of man does science make? The boy is not attracted. He says, 
I do not wish to be such a kind of man as my professor is. 

If the goal of a teacher is to stimulate, by his own example, en- 
thusiasm for his subject, Emerson s teachers of science seem not to have 
come very near this goal. It may be that Emerson was not a science- 
minded type. We may quote, therefore, a recent writer, who, besides 
his wide interests in all domains of human life, had a particular leaning 
toward science— H. G. Wells. He describes bis reaction to his physics 


228 



philosophy of science in the physics curriculum 


teachers in the Normal School of Science in London. He realizes exactly 
that what is wrong is not the ability of a particular teacher. “No man,” 
says Wells, "can be a good teacher when his subject becomes inexpli- 
cable.” He amplifies this statement: 

To a certain point it had all been plain sailing, a pretty science, with 
pretty subdivisions: optics, acoustics, electricity and magnetism, and so on. 
Up to that point, the time-honored terms which have been ciystaUized out 
in language about space, speed, force, and so forth sufficed to carry what I 
was learning. All went well in the customary space-time framework. Then 
things became difficult. I realize now that it wasn’t simply that neither 
Guthrie nor Boys was a good teacher . . . The truth, of which I had no 
inkling then, was that beyond what were (and are) the empirical practical 
truths of the conservation of energy, the indestructibility of matter and force, 
and so forth, hung an enigmatic fog. A material and experimental meta- 
physics was reached . . 

H. G. Wells indeed laid a finger on a very sore spot in our science 
teaching. The "subject has become inexplicable” because the teachers 
themselves did net have traming that pat the emphi^sis oa logical eon- 
sistency and satisfactory coherence between abstract law and sensory 
observation. Their upbringing was a purely technical one, with little 
regard for the role of science within our culture as a whole. Where 
physics teaching reached a domain beyond mechanical and electrical 
engineering, the subject was not satisfactorily explained because it was 
“inexplicable.” 

There is no doubt that the current interest in science has its origin 
not only in its technical application, but also in its bearing on our gen- 
eral world picture. Our generation is witness to the “relativity boom.” 
At the peak of this boom, Einstein’s theory was of hardly any use to 
the engineer. But, to express it perfunctorily, it changed our view 
about the “nature of time and space.” The statement that conciurently 
with a given event here a certain event is occurring on the planet 
Mars, was formerly taken to have a definite meaning. We know now 
that such a statement can only be asserted “relative to a certain sys- 
tem of reference.” This new doctrine appeared to add credit to the 

'■H. G. Wells, Experiment In Autobiography (New York: Macmillan, 1934), 
pp. 175” 


229 



modern science and its philosophy 


doctrine of the “relativity of truth” and seemed therefore to some 
people a “social danger,” since it might contribute to a disbelief in the 
“absolute values” of ethics. 

These problems have been haunting our generation for decades. 
But if we ask a trained physicist (not to speak of a graduate engineer) 
what his opinion is on these questions, we notice immediately that 
his training in physics has not provided him with any balanced judg- 
ment. The scientist will, as a matter of fact, often be more helpless than 
an intelligent reader of popular science magazines. We face the same 
situation if we ask our graduate in physics whether the theory of 
quanta has justified the belief in the freedom of will or whether it has 
made a contribution toward the reconciliation between science and 
religion, as has been maintained frequently, even by scientists and phi- 
losophers of high reputation. 

The great majority of these trained physicists— and by "majority” 
I mean more than 90 percent— will be unable to give any but the most 
superficial answer. And even this superficial answer will not be the 
result of their professional training, but the profit they have made from 
reading some popular articles in newspapers or periodicals. As a mat- 
ter of fact, most of them will not even be able to give a superficial an- 
swer, but will just say, “This is not my field, and that’s all there is to it.” 

The result of conventional science teaching has not been a critically 
minded type of scientist, but just the opposite. The longing for the 
integration of knowledge is very deeply rooted in the human mind. 
If it is not cultivated by the science teacher, it will look for other out- 
lets. The thirsty student takes his spiritual drink where it is offered to 
him. If he is lucky, he gets his information from popular magazines or 
science columns in newspapers. But it can be worse, and he may be- 
come a victim of people who interpret recent physics in the service of 
some pet ideology which has been, in quite a few cases, an anti- 
scientific ideology. The physical theories of the twentieth century have 
been interpreted, indeed, as having “abandoned rational thinking” 
in favor of— I don’t know exactly what, as I cannot imagine what al- 
ternative exists to rational thinking in the field of science. 

It is a fact which one may regret, but which is established em- 


230 



philosophy of science in the physics curriculum 


pirically with no less certainty than the law of gravitation, that the 
science student who has received tlie traditional purely technical in- 
struction in his field is extremely gullible when he is faced with 
pseudophilosophic and pseudoreligious interpretations that fill some- 
how the gap left by his science courses. I would even venture the 
statement that this gullibility increases in inverse proportion to the 
familiarity of the student with the conceptual analysis of science. I 
think that, statistically, the experimental physicist is more gullible 
than the theoretical physicist, and the graduate of an engineering 
school still more so. 

For our purpose, it is suflBcient to realize that the present type of 
science instruction does not enable the student to form even the faint- 
est judgment about the interpretation of recent physics as a part of a 
new world picture. This failure prevents the science graduate from 
playing in our cultural and public life the great part that is assigned 
to him by the ever-mounting technical importance of science to human 
society. It is obvious that the necessity of teaching twentieth-century 
physics has made conspicuous the shortcomings in our methods of 
science teaching. It would be erroneous, however, to think that the 
teaching of earlier, so-called classical, physics is not handicapped by 
the same kind of shortcoming. But, certainly, these deficiencies be- 
come the more conspicuous the more rapid the changes in physics 
that have to be presented in class. 

If we look, for example, at the treatment of the Copemican con- 
flict in an average textbook of science, we notice immediately that the 
presentation is far from satisfactory. In almost every case, we are told 
that according to the testimony of our senses the sun seems to move 
around the earth. Then we are instructed that Copernicus has taught us 
to distrust this testimony and to look for truth in our reasoning rather 
than in our immediate sense experience. This presentation is, to say 
the least, very misleading. Actually, our sense observation shows only 
that in the morning the distance between horizon and sun is increas- 
ing, but it does not tell us whether the sun is ascending or the horizon 
is descending. Starting from this fundamental mistake, the average text- 
book does not provide the student with an adequate picture of the 


231 



modern science and its philosophy 


historic fight of the Roman Church against the Copemican system. The 
student does not leam the way in which a powerful organization can 
oppose a doctrine established by science, and how this opposition can 
muster the support of reasonable and bright persons like the father of 
British empirical philosophy, Francis Bacon, who denounced the Coper- 
nican system as violating our common sense. 

By its failure to give an adequate presentation of this historic dis- 
pute our traditional physics teaching misses an opportunity to foster 
in the student an understanding of the relations between science, reli- 
gion and government which is so helpful for his adjustment in our 
modem social life. With a good understanding of the Copemican and 
similar conflicts, the student of science would have even an inside 
track in the understanding of social and political problems. He would 
be put at least on an equal level with the student of the humanities. 

Let us now proceed with our hearing: having listened to Emerson 
and Wells, let us question an intelligent college student who is not a 
concentrator in science but who has taken, say, a beginning course in 
chemistry or biology to round out his general education. We shall hear 
often of a deep dissatisfaction with these science courses as a contribu- 
tion to general education. While such a student may carry away from 
his courses in history or literature the impression that they have guided 
him into a wide-open field of human interest with an outlook in a great 
many new directions, he has had a very different reaction to his science 
courses. 

We may take our example from the chemistry classes, where, ac- 
cording to an old tradition, the general outlook is more neglected than 
in the other sciences. In such courses, where science teaching is at its 
worst from the educational angle, formulas like HgSO^ are thrown upon 
the head of the student, and he will really feel, as Emerson did, that 
these formulas are “notes in a pocketbook, which are valuable only to 
the owner”— who is, in this case, the professional chemist— but without 
any broad human interest. The general student will get the impres- 
sion that he sits in a show where formulas are tossed around in a kind of 
juggling game. This is, of course, a very inadequate impression, but 
one to which the student will inevitably be led by the purely tech- 



philosophy of science in the physics curriculum 

nical approach in the teaching. He will probably not be much im- 
pressed if he is told that this juggling leads eventually to interesting 
results, perhaps in the field of solid and liquid fuels. For “fuel” is, after 
all, no broad human interest either, but obviously a concern for some 
specialists. 

If, however, chemistry should be taught in a more human way, it 
would be clear that a formula like H2SO4 contains the history of man- 
kind, the evolution of the human mind in a much more impressive and 
certainly more condensed way than does all the history of the British 
kings. What a good teaching of science has to achieve on behalf of 
general education is to bring out the heroic mental efforts of man- 
kind, which are packed in the formula H2SO4 as in a nutshell, and to 
let the student live through the exciting historical and psychological 
experience that eventually found its abridged expression in such a 
formula. Then the “note in a pocketbook” will become a flamin g mani- 
festo to mankind. 

The foregoing discussion raises serious doubts as to the ability of the 
traditional method of science teaching in fitting science into the broader 
framework of human knowledge. For the scientist himself this state of 
affairs may be satisfactory or at least tolerable, provided that the 
student gets a good understanding of his own field, say physics. But 
I am afraid that the same shortcomings that have in the past been ob- 
stacles to fitting physics into a broader frame of reference may also 
be obstacles in the path leading to a satisfactory understanding of 
physics itself. 

In order to check this apprehension, we have only to put to a 
student of science the following simple question, which is, in a cer- 
tain way, a key question: What does it mean when you say that a 
geometric theorem— for example, the sum of the angles in a plane 
triangle is equal to two right angles— is “true”? The significance of this 
question was emphasized with the advent of non-Euclidean geometry. 
Even in a perfunctory treatment of non-Euclidean geometry the ques- 
tion will arise: What does it mean to say that our actual space is 
Euclidean or non-Euclidean? This question was regarded as very 


233 



modern science and its philosophy 


natural and pertinent by the founders of non-Euclidean geometry. It 
was, actually, a very common question when this new geometry was 
a novelty. But if, today, you examine a graduate in mathematics or 
physics, the majority will never have heard of such a question, and if 
you put it to them they will not understand its meaning without a very 
thorough interpretation. If we look into any current textbook of non- 
Euclidean geometry, which may have several hundred pages, hardly 
half a page will be devoted to the question of “truth,” and these few 
lines are mostly an attempt to dodge the question. In most cases, the 
student will not even learn such a simple thing as that mathematics can 
prove only statements of the type: if theorem A is true, then theorem 
B is also true; but it can never prove whether A is true or not. 

This failure of science teaching in the foundations of geometry has 
a devastating effect on the mind of the student. For if he does not 
grasp the exact relationship between mathematics and physics in this 
simplest example offered by geometry, later he will certainly be unable 
to understand precisely the relation between experimental confirmation 
and mathematical proof in the more involved domains of physics or, for 
that matter, in any exact science. This means that he is bound to mis- 
understand altogether the role of mathematical theory in physics. The 
failure of a satisfactory elucidation in such a fundamental matter has 
made it possible to find in a well-known textbook of college physics 
statements like this; Einstein proved “mathematically” that a material 
body cannot move with the speed of light. The student of physics 
is not even trained to get the instinctive feeling that no statement of 
physics can be proved mathematically, but that every "proof’ in physics 
consists only in deriving mathematically one physical fact from other 
statements about physical facts. 

By the failure to give a good account of the exact relation be- 
tween mathematical proof and experimental confirmation, our tradi- 
tional science teaching again misses an opportunity to teach the student 
a reasonable and scientific approach to all problems of human interest. 
For in all these fields the central problem is the relationship between 
sensory experience (often called fact finding), and the logical conclu- 
sions that can be drawn from it. The failure to grasp exactly the nature 



philosophy of science In the physics curriculum 

of this relationship accounts for the confused attitude of many people 
toward the complex problems by which they are faced in private and 
public life. 

The role of mathematics and physics in the understanding of geom- 
etry is perhaps the simplest example by which the student can learn 
how to discern the role of facts and logical conclusions in the involved 
problems of human relations. As a matter of fact, every problem of 
physics where mathematics is applied gives us such an example. A 
simple and typical example is provided by Newton’s laws of motion. 
The first law (law of inertia) and the second law look very simple, 
but they are a crucial issue for every instructor who teaches physics to 
beginners. Their apparent simplicity makes it easy to discover every 
confusion in the presentation, whereas in the treatment of an advanced 
subject of mechanics, such as the problem of three bodies in celestial 
mechanics, a little confusion can pass unnoticed. 

It is not an exaggeration to say that 90 percent of the textbooks 
of physics on the college level present the law of inertia in such a way 
that its meaning is obscure; it is formulated in words that are not ap- 
plicable to any actual situation in the physical world. We learn mostly 
that a body upon which no force is acting is “moving along a straight 
line.” But obviously the expression “moving along a straight line” has 
a physical meaning only if a system of reference is physically given, 
in which a straight line is fixed that serves as the model by which we 
judge whether a certain motion is rectilinear or not. But in the current 
textbooks of college physics one finds hardly even the perfunctory 
statements that the system of reference needed in the first law is the 
system of the so-called fixed stars. Hence, the statement of the law as 
it occurs in the textbooks has not even a vague meaning, for no method 
is described by which the validity of the law can be tested in a con- 
crete case. 

Since the method of testing remains obscure, one finds all imag- 
inable opinions regarding the possibility and method of testing the 
validity of the law of inertia. In some books one reads that the law is 
self-evident and does not need any empirical proof; other authors, 
however, say that it is confirmed by the most familiar experience of 


235 



modern science and its philosophy 


our daily life; for example, a book lies quietly ou a table if it is not 
taken off. Tills, again, is in contradiction to the assertion of some 
textbooks that the law of inertia is a hypothesis that cannot be proved 
by any experiment 

If one seeks an evaluation of this law of inertia as an achievement 
of the human mind, one finds not infrequently such statements as: it 
is ama zing that it took so many centuries to discover such a simple 
and trivial law. It does not occur to these authors that this law may, 
after all, not be so "simple” as they believe, if so many great scien- 
tists were unable to discover it earlier. Obviously, to call it “simple 
and self-evident” means not to understand its real significance and 
not to give to the students a correct presentation. 

While in the elementary textbooks the law of inertia is formulated 
in an elaborate but mostly meaningless way, many advanced textbooks 
take pride in minimizing the law. I even found a book in this category 
in which the law of inertia is called “self-explanatory.” The author 
himself fails to explain it, relying evidently on the ability of the law 
to explain itself. In these advanced books the law of inertia is treated 
as a special case of the second law: acceleration is proportional to 
force. If the force is zero, the acceleration obviously is zero too, and the 
body moves with constant velocity. But the only distinction between 
this and what the elementary books say is very often that in the ad- 
vanced book not the first, but the second law is formulated in a mean- 
ingless way. The result can be easily checked, if we ask any graduate 
student in physics what the word “acceleration” in the second law 
means. He will certainly know that the acceleration is the second 
derivative of the coordinates with respect to time; but when we ask 
him what is meant exactly by "coordinates,” he will say that they refer 
to a Cartesian coordinate system. But a motion of a physical body is 
described only if we specify its coordinates with respect to a co- 
ordinate system that consists of physical bodies at which we can point. 
Therefore, to give an “operational” definition of what is meant by 
“acceleration” in the second law, we have to describe a particular 
physical Cartesian system. This certainly must not be in fixed con- 
nection with our earth. In the first approximation, we can take the 


236 



philosophy of science in the physics curriculum 


body of fixed stars, our Milky Way, as such a Cartesian system. Then 
acceleration is, approximately, “acceleration with respect to our Milky 
Way.” Most students of physics will be at a loss if asked a simple 
question like this. Even a candidate in the Ph.D. examination will some- 
times answer that "acceleration” in the second law means “accelera- 
tion with respect to a Cartesian system the axes of which are fixed in 
space.” Many will not even have a vague apprehension that the ex- 
pression “fixed in space” may be meaningless in physics. But if one 
asks them a little more insistently, they will become confused and 
will guess that there is something rotten in their knowledge of the 
simplest laws of physics. 

This failure to describe the physical system of reference to which 
Newtons laws refer does not do any harm to the student who takes 
mechanics for its use in mechanical engineering. For in this restricted 
domain he can work on the assumption that our own earth is the 
system of reference that is meant in the definition of “acceleration.” 
In the extreme case, he will have to resort only to the system of fixed 
stars, as in computing the influence of the rotation of the earth on the 
orbit of a launched projectile. The engineer who has never heard of 
the difficult problem involved in the expression “acceleration” in New- 
ton’s law will by this failure not imperil the lives of people who use a 
bridge that was built according to his computations. 

Obviously this state of happy innocence becomes obnoxious if it is 
preserved in the treatment of cosmological problems. The fixed stars, 
of course, are not at rest with respect to one another and do not ex- 
actly form a rigid Cartesian system. Therefore, they cannot be sub- 
stituted in the Newtonian laws for the abstract term “Cartesian sys- 
tem.” Rather, we must say that the positions and velocities of the fixed 
stars, and even of the remote nebulas, determine in some way the 
system of reference that the Newtonians mean when they use the term 
“acceleration” without any specifications. This means that the motion 
of a rolling billiard ball on earth is physically influenced by the posi- 
tions and velocities of our Milky Way, and even by the galaxies that 
are millions of light years away from the earth. They determine the 
motion that we call “motion under no force.” By ignoring the influence 



modern science and its philosophy 


of the large but remote masses of the universe, the mysterious concept 
of “absolute space” was introduced into science. It remained there as a 
heritage of an ancient state of science which later was interpreted as 
metaphysics. 

Obviously, as in the case of geometry, by the failure to give a satis- 
factory presentation of Newton’s laws the teaching of physics misses a 
valuable opportunity. It could give to the student an example of the 
elimination of concepts that are leftovers from an earlier state of sci- 
ence and have survived in the disguise of metaphysical concepts. The 
“absolute space” and the “Cartesian system without physical back- 
ground” can be proved to be superfluous for any logical derivation of 
observable fact. The student understands easily that they have no 
legitimate status in science, and are only the source of an odd verbiage 
about "real motions” in contrast to “apparent motions.” 

In the sciences that deal with human behavior the situation is es- 
sentially the same. One makes use of expressions like “real freedom” 
in contrast to “pseudofreedom,” but the muddle around these expres- 
sions is very hard to disentangle. To learn how “real motion” was elim- 
inated from science is a very good example of how to proceed in other 
fields. Hence, it has an educational value that can hardly be over- 
estimated. It paves the way for a correct analysis in the social sci- 
ences. But we are not to restate this issue of value for general educa- 
tion in this section. Rather, we shall stress the point that the same short- 
comings which make the traditional presentation of Newton’s laws 
unfit for general education, bar also its use for the understanding of 
physics itself, particularly of recent physics. 

It is a fact which can hardly be overlooked that in the average 
college curriculum little attention has been paid to the theory of 
relativity compared with the great interest which this theory has met 
among the general public, and compared with the bearing of this 
doctrine on atomic and nuclear physics, not to speak of its key role in 
the logic of science. When, for example, the theory of the electro- 
magnetic field is treated in the regular college curriculum, there are 
many teachers who skip the relation to the theory of relativity or treat 
it perfunctorily. They will tell the student that “relativity” is an obscure 



philosophy of science in the physics curriculum 

theory which the student would hardly understand. By treating the 
subject in tliis perfunctory way lliey certainly contribute to the 
obscurity that may prevail in the mind of the student. He becomes 
convinced that the theory of relativity is fundamentally different from 
“sound” physics, such as Newtonian mechanics or Maxwell’s electro- 
magnetic field. He is likely to believe that modem theories are less 
logical and less closely related to the observed facts. The origin of this 
attitude must be found, certainly, in the insufficient training of teachers 
in the field of relativity. As a matter of fact, there is even a shortage of 
appropriate textbooks. There are good textbooks on relativity for ad- 
vanced students and there are fairly good books for laymen, but a book 
on this subject for the undergraduate hardly exists. 

It is easy to put the finger on the gap in the training of college 
teachers which produces their antipathy to a thorough treatment of 
relativity theory in the regular physics courses. The lack of precision 
in the foundations of geometry and mechanics at which we pointed 
does no harm in mechanical engineering, but becomes a major danger 
when we have to teach relativity. To the mechanical engineer the 
whole problem of the system of reference is trivial. He can get along 
splendidly with our good old earth as such a system. But when we 
go in for more general problems, such as motion with great speed, 
propagation of light, and so forth, we have to keep in mind constantly 
that these phenomena can be described with respect to different sys- 
tems of reference. For some particular changes of the system of ref- 
erence the laws of motion remain unaltered. This fact turns out to be 
the most important property of these laws, and is the very basis of the 
theory of relativity. 

If the teacher fails to stress the role of the system of reference in 
ordinary mechanics, the student will later fail to grasp the gist of the 
theory of relativity. Moreover, if words like “absolute space” or "ab- 
solute rest in space” are not eliminated by adequate teaching of me- 
chanics, the student will have trouble eliminating the concepts of 
“absolute length” or “absolute time” in relativity. Frequently, the teach- 
ing of traditional college mechanics has not been used to give to the 
student a correct understanding of the relation between mathematics 


239 



modern science and its philosophy 


and observation in science, but instead the metaphysical formulation of 
Newton’s laws is hammered into the student’s brain, as is done in most 
textbooks. Then the study of the theory of relativity will be a hard 
job for instructors and students. They will easily wind up in the pit- 
falls of metaphysical prejudices which are mostly covered by a thin 
stratum of common sense. They will believe that expressions like “ab- 
solute simultaneity” or “absolute length of a time interval” are im- 
mediately understandable to common sense. The result is, of course, 
that these students, on becoming instructors in physics, will never be 
able to get rid of the opinion that the theory of relativity is somehow 
contrary to common sense, or, at least, not as agreeable to it as tra- 
ditional physics has been. 

This situation is still worse in the approach to the quantum the- 
ory. As this theory is badly needed to organize experimental results in 
atomic physics, it would be hard to eliminate the quantum theory 
from the regular physics curriculum. But the teachers who do not feel 
on firm ground in the foundations of geometry and mechanics will 
teach the quantum theory as a collection of useful recipes, “quantiza- 
tion laws,” which often take the form of “prohibitions.” They will dodge 
as much as possible the new general laws that have replaced the New- 
tonian science. The student will get little of Heisenberg’s principle of 
indeterminacy or Bohr’s principle of complementarity, which are the 
most relevant points in quantum theory for our general world picture. 
And if these points are touched, the same thing will happen as in the 
ordinary treatment of relativity. The student will be left with the 
impression that the new physics is somehow obscure and even “irra- 
tional.” This perfunctory presentation of the new principles is, in some 
respects, even worse than the practice of restricting the teaching pre- 
cisely to the formulas that describe actual observations. For the feel- 
ing that science has to be satisfied with an obscure theory makes the 
student an easy victim of quacks who exploit modem science in their 
endeavor to prove that we have to surrender our reason to a kind of 
blind instinct. This means in practice that we have to surrender our 
good judgment to the control of people who are bold enough to pre- 
tend that they are in possession of that intuitive instinct. 


240 



philosophy of science in Hie physics curriculum 


To summarize these remarks: if the teacher of science takes it easy 
in the presentation of the foundations of geometry and mechanics, the 
consequences of this carelessness will leave imprints on the minds of 
the student that will make it very hard to train him later as a teacher 
of modem physics. 

H. G. Wells, in his criticism of the science teaching he went 
through, says very correctly: 

The science of physics is even more tantalizing today above the level of 
elementary introduction (optics, acoustics, etc.). Brilliant investigators 
rocket o£E into mathematical pyrotechnics and return to common sense with 
statements that are, according to the legitimate meanings of words, non- 
sensical. Ordinary language ought not to be misused in tliis way. 

This statement gives a good diagnosis of the disease by describing 
it as basically a “semantic maladjustment,” as the school of general 
semantics would call it today. “Clearly,” Wells continues, “these mathe- 
matical physicists have not made the real words yet, the necessary 
words that they can transmit a meaning with and make the basis of 
fresh advances.” As we shall see in the next section, these words even 
hint at the remedy of the illness. 

We have shown that our traditional way of teaching science is re- 
sponsible for two deficiencies in the education of our students. First, 
not all instmctors in science are up to the role that science, particularly 
physics, has to play in our general cultural life. Second, the twentieth- 
century theories (relativity and quantum theory) are still an obscure 
domain outside the well-illuminated field of common-sense physics. 
Uneasiness has not been removed by our usual teaching of physics. 

If we now put the question of how to improve this situation, we 
know from the preceding discussion the sources from which the failure 
of our traditional teaching has originated. We know how it has hap- 
pened that our science instruction is not successful in fitting science 
into our cultural life, and the modem theories into the frame of 
common-sense physics. 

We discovered the main source in the failure to teach the students 
a clear distinction between the observation of facts and the drawing of 


241 



modern science and its philosophy 


logical conclusions. David Hume, the father of empiricist philosophy, 
has urged us to “surrender to the flames” all books that contain neither 
observed facts nor mathematical conclusions. This recommendation 
has become the basis of all attempts to make science a coherent and 
empirically confirmed system of knowledge. Accordingly, we have to 
teach our students how to eliminate all statements that are neither 
propositions describing the result of an observation, nor propositions 
that are part of a logical conclusion. 

Therefore, we have to insist, above all, that the student leam to 
analyze the statements of his science in such a way that it becomes 
obvious what is a statement of observation and what is a logical con- 
clusion. Many statements made in the traditional presentations of 
physics, particularly of recent theories, do not belong to either of the 
two types. 

If we have carried out this analysis, it is convenient to present the 
result in such a way that from the wording of every statement it imme- 
diately becomes clear what its logical status is. For, in the traditional 
presentation of science, we are often unable to recognize whether a 
given statement is meant to be a statement about observed facts, or a 
hypothesis about facts that will be observed, or a statement that merely 
introduces new words or rules for construction of propositions. This 
distinction, however, is necessary. 

If we try to single out the purely logical statements, we find them 
in every science. In geometry, for instance, a great part of scientific 
work consists in deriving conclusions from a given set of propositions 
called axioms. In drawing these conclusions, we do not need to know 
what the words in the axioms mean in the physical world. If the 
“points” and “straight lines” and “triangles” have the properties that 
are ascribed to them in the axioms, they also have the properties that 
are ascribed to them in the theorems, which we obtain by logical con- 
clusions from the axioms. In the same way, in mechanics, if we start 
from Newton’s equations, ma — f and a = d“x/dt®, we can derive, by 
logical conclusion, a great many theorems. We find that if the force is 
zero, the coordinate x is a linear function of time, and if the force is con- 
stant, the coordinate x is a quadratic function of time, and so on. These 


242 



philosophy of science in Ae physics curriculum 


statements are true whatever the meaning of “force f or “coordinate at” 
or “time t“ may be in the physical world. 

We obtain in this way purely logical statements. One occasionally 
calls them “tautological statements” because they tell us only that one 
and the same assertion can be expressed in diflFerent ways. If ma = f 
and f = 0, it follows that x is a linear function of t. But from this con- 
clusion we do not learn anything about the physical world; we learn 
only to say one thing in different ways. By this perfunctory hint we 
learn already that more than a few failures in our customary science 
teaching have originated in the failure to recognize clearly the tauto- 
logical character of these statements. 

The world of physical experiments— briefly, the “physical world” 
—is described by statements about the procedure and results of physical 
measurements. They boil down, eventually, to statements like "a 
certain colored spot covers a second colored spot,” or statements about 
a “pointer reading,” as Eddington calls it. They consist only of words 
that are familiar to everybody. The description of physical experi- 
ments does not contain any element that is not used in describing our 
breakfast, and is therefore understandable to everybody who is trained 
to use the English language in his daily life. These statements have, 
obviously, a certain vagueness, and they are never completely un- 
ambiguous. 

These “observational statements” are the second type of statement 
that we can single out in any science. In geometry, for example, we 
show how to measure the angles of a triangle and can And that the 
sum of the results is approximately 180°. In mechanics we measure the 
distance of a body from the floor of a room, and can find that in the 
case of a launched ball the distance from the floor is approximately 
a quadratic function of the angle traversed by the hand of a clock. 

These two types of statements, the logical and the observational, 
contain different types of word. In the first one, we have symbols like 
“points” or “forces” or “coordinates,” which have no meaning at all in 
the physical world, while the second type contains words of our 
‘Tdtchen language,” which denote familiar things of our environment. 

Customarily, the student learns in science classes that by observa- 


243 



modern science and its philosophy 


tional statements it is possible to confirm or refute the axioms of geom- 
etry or laws of mechanics. But, obviously, this way of speaking is some- 
what superficial and logically incorrect. For the “observational 
statements” contain words of everyday life like “green,” “hard,” “over- 
lapping spots,” and so forth, while the “axioms of geometry and me- 
chanics” contain symbols like “point,” “circle,” “acceleration,” “force.” 
A statement of the second type can never be confirmed or refuted 
^ by a statement of the first type, because it does not contain the same 
terms. 

If science consisted only of principles and observations, its validity 
could not be tested by scientific methods. It is, therefore, of the utmost 
importance to teach the student that, besides these two types of state- 
ment, science has to make use of a third type by which the language 
of the abstract principle can be translated into the language of obser- 
vation. If the principles contain, for example, the word “length,” one 
has to describe the physical operations by which the length can be 
measured. As these operations can be formulated in the language of 
everyday life, of observational statements, the description of the oper- 
ation translates the abstract term “length” into terms of observational 
language. 

These translating sentences have been studied with particular em- 
phasis and care by F. W. Bridgman, and are called, according to his 
suggestion, “operational definitions.” By this definition, an abstract 
term like ‘length” acquires “operational meaning.” When these “oper- 
ational definitions” are introduced into the abstract principles, the lat- ' 
ter become “observational statements.” Then we can check by actual 
experiments whether these observational statements, derived from the 
principle, check with the observational statements that describe our 
direct physical experiments. It is necessary to understand that only in 
this indirect way can the principles of science be checked by experience. 

Nowadays, the term “semantics” is often used to denote the study 
of the relation between words and meaning. Hence, the sentences by 
which symbols are coordinated with observable things are called 
“semantic rules.” They occur in die writings on the logic of science 
under different names. F. S. C. Northrop uses them as “epistemological 


244 



philosophy of science in the physics curriculum 

correlations,” H. Reichenbach as “rules of coordination,” and so on. 

This “logico-empirical analysis,” or “semantic analysis,” is a general 
method of “making our ideas clear.” It is an eflBcient method for com- 
munication of thoughts and for influencing people. This analysis is the 
chief subject that we have to teach to science students, in order to fill 
all the gaps left by traditional science teaching. 

If the student were taught this method effectively, he would be 
able to present and to understand recent physical theories— like rela- 
tivity and the quantum theory— as precisely as classical physics, if— 
and this is the great “if”— if he understands ordinary geometry and 
mechanics precisely. Therefore, as charity should start at home, so 
semantic analysis should start in our own scientific back yard, at the 
foundations of our familiar geometry and mechanics. The application 
of logico-empirical analysis was never presented so clearly and con- 
cisely as in Einstein’s paper, “Geometry and Experience.” He sum- 
marizes the results by saying; As far as geometry is certain, it does 
not tell us anything about the physical world, and as far as it makes 
statements that can be tested by experience it is not certain— that is 
to say, not more certain than any statement of physical science. By 
operational definitions "straight lines” can be translated into “light 
rays” or “edges of rigid bodies.” Then the theorems of geometry be- . 
come statements of optics or mechanics. Without these definitions, all 
(porems of geometry are tautological propositions about symbols, 
ticsioreover, if we apply this type of analysis to mechanics, we shall 
1 step Uv ask: what is the operational meaning of “acceleration”? If 
we wfiil aii< translate this symbol into an expression consisting of ob- 
servational terms, we have to describe the physical system of reference 
to which the coordinate x refers. 

If we apply the same method to the theory of relativity, this doc- 
trine will lose its isolated and obscure character. It will be obvious that 
“length” or “simultaneity” have to be defined by the description of 
physical operations by which these quantities can actually be meas- 
ured. If yardsticks and clocks are affected by motion, it is equally 
obvious that the operational definitions of spatial or temporal ‘length” 
have to contain the velocity of the yardstick or clock relative to a physi- 


245 



modem science and its philosophy 


cal system of reference. This means that only “relative length” has an 
operational meaning, not "absolute length.” 

The reason why the theory of relativity has appeared so obscure 
to the physicists will then be very clearly understood by the student. 
Every alteration of the laws of physics seems understandable to com- 
mon sense, if the semantic rules are unaltered. This would be the case 
if we replaced the wave theory of light by the corpuscular theory, pro- 
vided both theories remain within the frame of Newtonian mechanics. 
But the experiments on bodies moving with great speed suggest the 
introduction of new semantic rules. Such a change is much more funda- 
mental and will appear obscure to the student of physics, if he is not, 
from the start, taught the role of these semantic rules in every do- 
main of science, beginning with his high school plane geometry. 

Bridgman rightly stresses the point that for a student who has 
studied relativity successfully, there is no need for a subsequent intro- 
duction to logico-empirical analysis. He could not understand relativity 
without it. The instruction in the two theories is inseparable. 

In a similar way, if we apply this analysis to the quantum theory, 
no student would fall for the talk that there is some “irrational” or 
“organismic” aspect in the quantum theory. He would understand that 
the very meaning of Bohr’s principle of complementarity can be defined 
as the introduction of new semantic rules into the language of physics. 
As the alterations required by these rules are much more radical thct 
those required by the relativity theory, the confusion of languag per- 
been tremendous. As our traditional teaching fails often to me lat- . 
correct logico-empirical analysis, a great many author. ■ by and 
philosophers alike— have failed to recognize that we have to ao with a 
radical alteration of semantic rules. Many of them have even been 
caught by a widespread propaganda by which the new physics of the 
twentieth century has been exploited in die service of a campaign 
against the spirit of science, in favor of a confused “organicism,” “spir- 
itualism” or “irrationalism.” 

Clearly, by learning the application of semantic analysis, the stu- 
dent will not only profit from a better understanding of his own science, 
but also gain a more correct appreciation of the role his science plays 



philosophy of science in the physics curriculum 

in the frame of human culture in general. The close relationship of 
semantic analysis to modem physical theories on the one hand, and to 
the understanding of our general social and cultural life on the other 
hand, has perhaps nowhere been stressed so bluntly and sharply as by 
Bridgman, in a paper published in 1945.' He discerns the technical and 
the semantic significance of twentieth-century science, when he says: 

The second aspect of the modem epoch in science is, I believe, of in- 
comparably greater significance ... He [the physicist] has come out with 
what amounts to a new technique of analysis, of great power and unexpect- 
edly wide range of applicability. The new technique is applicable to all ques- 
tions of meaning . . . The potentialities of the new technique, when applied 
to domains outside the present application of science, may be glimpsed by 
contemplating the confusion which now reigns ... It is becoming evident 
to an increasing number of people that an important part of our difiBculties 
in analyzing and conveying meaning is of verbal origin . . . There are 
popular discussions of semantics. 

To appreciate this popular trend toward semantics, we need only 
look at such books as Hayakawa’s Language in Action,^ which was a 
best seller. In ETC; A Review of General Semantics, edited by Haya- 
kawa, semantics is applied to the problem of race relations, to labor 
problems, to the psychology of fear, and so forth. 

Our point has been tliat an important step toward improvement of 
our science teaching could be taken by consistent application of seman- 
tics or logico-empirical analysis. I do not believe, however, that this 
step would be sufficient. Even if such an analysis is carried out in a 
careful and competent way, there still remains much to be done if we 
want to bring out all the educational value that is inherent in science. 

Once when I explained to a class the second law of Newton by the 
method of logico-empirical analysis, stating precisely the operational 
meaning of the symbols m, a and f, a student asked: Is this really every- 
thing contained in the law? Has die intuitive conception of a “force 
acting upon a body,” which has played such a decisive role in the his- 

*P. W. Bridgman, Yale Review 34, 444 (1945). 

‘ S. I. Hayakawa (Harcourt, Brace, 1941}; also see W. Johnson, People in 
Quandaries (Harper, 1946). 


247 



modern science and its philosophy 


tory of our thought, completely evaporated into the thin air of logical 
rules and physical measurements? As another example, everybody vv^ho 
has ever participated in discussions about the principles of physics 
occasionally encounters the person who becomes excited when the 
question is raised whether the concept of force can be eliminated from 
physics. At this point an emotional element enters the debate. The 
heat of the argument had its origin certainly not in science itself, but 
in the stretch of history of the human mind, which is now packed 
into the word “force.” In the time when the Nazi party and its philoso- 
phy flourished, we could frequently find in German writing the view 
that the concept of “force” is somehow connected with the Nordic race. 
All people who wanted to eliminate it from physics were branded as 
enemies of the Nordic race. On the other hand, those who plead for the 
elimination of the concept of “force” regard it as a remainder of an 
animistic and anthropomorphic world picture, which even has a touch 
of spiritualism and superstition. 

We learn from this example that, in order to understand the role 
that the concepts and principles of science have played in our cultural 
life, we need, in addition to the semantic, an analysis of a different 
type. We have to learn not only the operational meaning of symbols 
like “force” and “mass,” but also how it has come about that just these 
symbols were chosen. Obviously the choice has not been unambigu- 
ously determined by their abihty to form a basis for the derivation of 
observable facts. The symbols also have a life of their own. If we go in 
for this kind of research, we find that extrascientiflc reasons have been 
^ largely responsible for the predilection in favor of or the aversion to 
some symbols. We learned it from the example of “force.” A careful 
and thorough investigation has shown that psychologic reasons have 
played their part, as have habits carried since childhood, and even 
wishes emerging from the subconscious. 

Besides these factors of individual psychology, the religious, social 
and political trends of the period have been responsible for the pre- 
dilections and aversions of which we spoke. What we need, therefore, 
for an all-round understanding of science is an analysis of the psycho- 
logic and sociologic factors that went into the determination of our 


248 



philosophy of science in the physics curriculum 

scientific symbolism. We may speak, briefly, of a “socio-psychologic” 
analysis. We can summarize and say that this second type of analysis 
has to be added to the semantic analysis.^ 

Both types of analysis are parts of what one calls in more or less 
popular discourse “philosophy of science.” Thus we may say that the 
remedy for the afore-described shortcomings in our traditional way of 
teaching science is to give more attention to the philosophy of science. 
More precisely, we mean by it logico-empirical and socio-psychologic 
analysis, or, in the terminology of the school of thought known as 
logical empiricism: we need semantic analysis and “pragmatic anal- 
ysis.” 

We have now to attack the more practical question of how the 
desirable instruction in the philosophy of science should and could be 
given to the students of science. 

In my opinion, the best solution is that the elementary science 
courses themselves should contain a good deal of the philosophy of 
science. This solution presupposes, of course, that there are a sufiBcient 
number of science teachers available who have adequate training in 
the philosophy of science. Speaking now in terms of what should be, 
rather than in terms of what is, I believe that no teacher should give an 
introductory course in science if he lacks training in semantics and the 
history of science. I realize that this solution is not feasible today. 
Hence, my second choice is to introduce courses in the philosophy of 
science which are to be obligatory for the students of science, at least 
for those who are later to become teachers of science. In a certain way 
every scientist is a teacher of science in his environment, whether he 
works as an engineer, a physician, or even a pharmacist. 

Courses in the philosophy of science for the benefit of science stu- 
dents have been given rarely; hence there is no coherent tradition. We 
are faced with the problem of building up such a tradition. There are 
two reefs that have to be avoided. The greatest weakness of the at- 
tempts which have been made occasionally to give such courses has 

* Jean Piaget, Les Troit Conditions cTtme iplstdmologle scientipque. Analysis 
(Milan, 1940), Vol. 1, No. 8, p. 25. 

J 


249 



modern science and its philosophy 


been the lack of any definite conviction on the part of the professor. 
We must never forget that the word "professor” is derived from 
“profess,” and this holds true, particularly, for a teacher of philosophy. 
His work with the students should be a “profession” in both meanings 
of this word. Scientists who have been teaching the philosophy of 
science have mostly oflFered a kind of incoherent digest of philosophic 
opinions. The choice has been mostly determined by the chance ac- 
quaintances of the teacher. The students have remained unimpressed 
by such courses or books. We meet this eclectic attitude even in the 
writings of such excellent scientists as Jeans or Planck. I remember the 
lectures of a great physicist, Boltzmann, on the philosophy of physics, 
which I attended as a student. Despite the personal greatness of the 
lecturer, the effect of the course was slight, because of a lack of a 
coherent approach. We can notice, on the other hand, that scientists 
who built their books around a central idea have shaped the minds of 
science students for decades. I mention, just as examples, Mach, 
Poincare and Bridgman. 

Although a coherent approach is the first requisite for successful 
instruction, the second reef on which our course could be wrecked is a 
narrow-minded indoctrination. I remember that an intelligent girl 
student told me once: “It is a very unpleasant feeling to have a certain 
doctrine drilled into your brain without being provided with an out- 
look into the open field of widely divergent opinions.” These two 
requirements of a coherent approach and of an open field seem to 
exclude each other. But this is true only superficially. If the socio- 
psychologic approach is added, it becomes clear that every opinion 
which has been advocated in history has its legitimate place within a 
coherent presentation of science, based on logico-empirical and socio- 
psychological analysis. 

Our problem now is, practically, to build a course in which the 
student gets training in both types of analysis— a course that not only 
fills the gap left by the traditional science teaching, but also leads into 
the wide field of controversial ideas for bridging this gap. 

For six years I have given a course at Harvard with this goal in 
mind. I have tried to improve it from year to year according to my 



philosophy of science in the physics curricuium 


increasing experience with students. I will describe the content of this 
particular course because of my familiarity with it, and not because I 
believe that it necessarily represents the best possible course. The 
course takes two terms. Actually, I give two half-courses which are 
designed for students of diflFerent backgrounds. But in this description 
I am going to integrate both half -courses into one coherent presenta- 
tion. 

My starting point has been the traditional distinction between 
“scientific truth” and “philosophic truth.” Specifically, I have referred 
to the formulation that the great medieval philosopher, Thomas 
Aquinas, gave to this distinction. According to him, there are two 
criteria of truth: the first requires that a proposition can be logically 
derived from “evident principles” (philosophic truth); the second 
requires that the logical sequences of the proposition in question can 
be confinned by experience (scientific truth). Obviously, if the truth 
of the principles themselves can be tested only by comparing their 
logical consequences with experience, the two criteria merge and we 
have only one criterion, namely, the scientific one. Then the traditional 
distinction disappears. This is the view that is called "empiricism,” or 
“positivism.” To assume the distinction obviously means to believe in 
principles the truth of which can be checked by extrascientific methods. 
In modem science this distinction is no longer upheld as an absolute 
distinction, but it has always played a certain role in a weakened and 
“relativized” form. We still say that there are propositions of physics 
which can be “proved” if we use the traditional language of physics 
textbooks. This means that they can be derived from general prin- 
ciples (like conservation of energy) which we hold to be true with a 
high degree of plausibility. Distinguished from these are propositions 
that are records of direct observation and cannot be “proved.” 

After these introductory remarks, a historical survey is given of the 
principles that have actually been used by the scientists as the basis 
from which the propositions of physics have been logically derived. 
There is one period in history that we can study from the origin to the 
end. This is the period of mechanistic physics, which lasted approxi- 
mately from 1600 to 1900. Its basic principles were Newton s laws of 


251 



modern science and its philosophy 


motion. A proposition of physics was regarded as “explained,” or 
“understood,” or “proved to be true,” if it could be logically derived 
from Newton’s laws of motion. But the question arose, of course, 
whether Newton’s laws, themselves, were “evident principles.” The 
belief that they are arises very often among students of physics as a 
result of the perfunctory way in which these laws are presented in the 
traditional textbooks. The only effective way to destroy this belief is 
to give a thorough presentation of the period before 1600, which pre- 
ceded the era of mechanistic physics. In my course, medieval and 
Aristotelian physics is presented and analyzed carefully. I call it 
“organismic physics,” because the basic principle is the analogy be- 
tween physical phenomena and the behavior of organisms. 

If the student has become aware that there was a time when reason- 
able people did not believe in Newton’s mechanics, it will not con- 
travene his “common sense” to learn that after 1900 a period could 
start in which these classical laws of motion were modified radically. 
Having become familiar with a conception of physics that prevailed 
before the period of mechanics, the student will look at the disintegra- 
tion of mechanistic physics around 1900 as a phenomenon analogous 
to the disintegration of organismic physics around 1600. In both cases, 
before the actual revolution took place under the impact of new dis- 
coveries of facts, the belief in the certainty of the ruling principles was 
shaken from within by logically minded critics. In medieval physics 
this role was played by the school of nominalism (Occam’s razor), 
while the criticism of mechanistic physics came from the “positivists” 
of the last quarter of the nineteenth century, men like Stallo in America 
and Mach in central Europe. 

The rise of twentieth-century physics, of relativity and the quantum 
theory, was closely connected with a new view of the basic principles. 
It was no longer taken for granted ffiat the principles from which the 
facts had to be derived should contain a specific analogy, either to an 
organism or to a mechanism. Nothing was required except that the ^ 
observed phenomena could be derived from the principles in a con- » 
sistent way and as simply as possible. The words and symbols that 
occurred in the principles, and the way these were coimected, could 



philosophy of science in the physics curriculum 

be invented according to their fitness as bases for deriving the phe- 
nomena discovered by the experimental physicist. 

This means that the symbols used in the principles had in them- 
selves no meaning beyond their value for the derivation of facts. From 
this introduction of abstract symbols arose the obvious need for in- 
cluding in the theory prescriptions for relating these symbols to sense 
observations. The necessity and the importance of these sentences were 
particularly emphasized and elaborated by Bridgman in his require- 
ment of “operational definitions.” This new aspect of physics, which 
dismissed any traditional analogy and stressed only the criterion of 
empirical confirmation and logical coherence, may be called “logico- 
empirical physics.” 

Under this name I refer in the course to the physics of the twentieth 
century, and dismiss by this terminology aU misleading metaphysical 
interpretations of relativity and quantum theory which we owe to phi- 
losophers, sociologists, theologians and— I am sorry to say— also to 
some physicists. 

After this historical framework has been established, I proceed to 
the application of logico-empirical analysis to the most relevant parts 
of physics. The start is made with geometry, where the situation is 
easier to explain than in any other field. I believe that the logico- 
empirical analysis of geometry provides the student of science with the 
best introduction to the philosophy of science. I would even venture 
to say that, if there is little time available, the familiarity of the science 
student with the analysis of geometry would suffice to fill, in a fairly 
satisfactory way, the gap left by traditional mathematics and physics 
teaching. By this analysis the student will learn that mathematics can- 
not prove any facts, but can only derive facts from other known facts. 
He will learn that there are two fields of science that we call 
“geometry.” “Mathematical geometry” has the previously described 
tautological character and cannot teach us anything about the “nature 
of space.” He will learn that only by the introduction of operational 
definitions— for example, “a straight line is a light ray”— can theorems 
of geometry become statements about facts and be confirmed experi- 
mentally like any law of physics. He will learn that no geometric 


253 



modern science and its philosophy 


theorem by itself can be confirmed by experience, but only a geometric 
theorem— such as the one about the sum of angles in a triangle— plus 
the operational definitions, of straight lines, and so on. He will learn 
that statements like "the axioms are self-evident,” or even “they are 
true,” are meaningless if no operational meaning is given them. 

I would even venture to say that a thorough semantic analysis of 
geometry contributes more to the intellectual outlook of the student 
than a superficial philosophic interpretation of the whole field of 
physics. By this analysis of geometry the student will be introduced to 
the new logical techniques in modem physics. Bridgman says of these 
techniques that their impact on the pattern of our culture will eventu- 
ally be greater than that of the technical consequences of modern 
physics, including atomic energy. The logico-empirical analysis of 
geometry will, moreover, enable the student to form a sound judgment 
about the "tmth” of non- Euclidean geometry, and in this way fill a gap 
that yawns in most textbooks. The problem of “tmtli” is usually dodged 
in the presentation of non-Euclidean geometry, so that this part of 
mathematics has remained obscure to most students. 

Then the course passes to the analysis of Newtonian mechanics. 
From the beginning, the operational meaning of terms is stressed. This 
means, above all, that the role of the system of reference emerges 
clearly. Newton’s laws are presented not as self-evident, but as para- 
doxical, for so they appeared to Newton’s contemporaries. 

We proceed then to the mechanical theory of light. When inter- 
action of light propagation with the motion of large material bodies 
occurs, and both phenomena are assumed to follow Newton’s laws of 
motion, one can predict phenomena that are in contradiction to ob- 
served facts, as in Michelson’s experiment. 

This contradiction is the origin of the restricted theory of relativity, 
which is analyzed thoroughly with the emphasis on the logico-empirical 
aspect and on the place of operational definitions. Then all those mis- 
understandings disappear that have made this theory “obscure” and 
“contradictory” to common sense. It becomes obvious that the role of 
the observer is not different from his role in Newton’s physics. No sub- 
jective element occurs in physics and, hence, there is no argument in 



philosophy of science in the physics curriculum 

favor of idealistic philosophy. The revolutionary changes brought about 
by relativity consist, first, of new observational facts, such as the change ^ 
of the rate of clocks by motion, and second, in the suggestion which 
Einstein made to alter the language in such a way that these new 
phenomena can be described in a simple and practical way. The rela- . 
tivily of time, for example, is no advance in metaphysics, but it is in I 
physics and semantics. 

In the same way the quantum theory is analyzed. The famous rela- 
tion of indeterminacy has nothing whatever to do with the introduc- 
tion of a subjective or spiritual factor into physics. It is actually a sug- 
gestion for introducing a language that best fits the facts. Since there is 
no law of physics that incorporates the expression, “the position and 
velocity of a small particle at one and the same instant,” there is no 
reason to introduce such an expression into the language of physics. 
Bohr’s principle of complementarity suggests that our physical world 
be described in a complementary language. 

The course then proceeds to a logico-empirical analysis of the con- 
cepts of causality, determinism, indeterminism, and chance along the 
same lines. Eventually the concepts of mass and energy are analyzed. 

After this description of the way in which semantic analysis is 
actually applied to physics, the course turns to problems of what has 
traditionally been called “philosophy.” The first, historical part was a 
survey of the way in which principles of science have developed from 
the organismic to the mechanistic physics, and further to the logico- 
empirical stage. In this stage there was no demand for a specific type 
of principle or a specific form of words in the principles. The wording 
of the principles became irrelevant from the viewpoint of science itself. 
But there have been extrascientific reasons for insisting on principles 
of a certain type. These reasons include mere sluggishness in changing 
old principles or, very often, theological, ethical, or political factors. It 
has been argued that the principles of science have to be in agreement 
with the prevailing principles in those extrascientific fields. We re- 
member the Thomistic distinction between philosophic and scientific 
truth. This distinction is based on the belief that one can assume some 
principles for reasons which are not in agreement with their conse- 



modern science and its philosophy 


quences as experienced. This means that one assumes these principles 
for extrascientific reasons. Every system of thought that has this belief 
as its basis is a metaphysical system. Since in these systems the prin- 
ciples are closely connected with the religious, moral and political 
predilections of a certain period, the study of the metaphysical founda- 
* tions of science is important for the understanding of the position of 
science in our cultural life. Therefore, the next part of the course is 
devoted to the metaphysical and antimetaphysical interpretation of 
science. 

The most important point, as it seems to me, is to get the student to 
understand the extremes. I take pains to present an adequate concep- 
tion of “straight metaphysics” and, at the other extreme, “straight 
positivism,” which bluntly says that there is no principle except those 
which can be confirmed by the agreement of their consequences with 
experience. 

As the example of straight metaphysics, the Thomistic philosophy 
is presented at length as far as it serves as a basis of science. The stu- 
dents are encouraged to read a textbook of “cosmology” that is cur- 
rently used in Catholic colleges and contains the Thomistic founda- 
tions of physics. Once I asked a teacher at a Jesuit college, who was 
taking my course, to explain to the students the authentic teaching of 
Thomistic philosophy. 

Then I proceed to the other extreme, which denies the distinction 
between the two kinds of truth and recognizes only scientific truth. 
I give a short historical survey starting with David Hume, followed by 
a short account of A. Comte’s “positive philosophy” and a more elabo- 
rate discussion of positivism at the end of the nineteenth century; of the 
Americans, Stallo and C. S. Pierce; of the Europeans, Mach and Poin- 
care. Passing to the twentieth century, we discuss the “operationism” 
of Bridgman and its relation to the pragmatism of William James and 
Dewey. Bridgman’s emphasis on the role of language leads us to the last 
phase of positivism, the logical positivism of Wittgenstein, Carnap, and 
others. 

After the presentation of straight metaphysics and positivism, we 
^discuss some attempts to produce a reconciliation between the two 


256 



philosophy of science in the physics curriculum 


views. There are scientists and philosophers who present science gen- 
erally in the positivistic manner, but reserve a jeparate compartment 
for metaphysics. Among them are Spencer, Fiske and P. Duhem. Sys- 
tematic attempts at reconciliation have been made on the basis of ideal- 
ism and materialism. The most powerful idealistic attempt is the school 
of thought founded by the German philosopher, Kant. Proponents of 
this school claim that there are self-evident principles, but that they are 
actually statements about our own minds. By making statements about 
geometry, we actually make statements about our own way of de- 
scribing nature. Therefore, these statements are, in a certain sense, 
checked by experience, but not by external experience. We observe om:- 
selves and believe that we observe the external world. An outstanding 
representative among the scientists of recent days of this Kantian com- 
promise is the British astronomer, Eddington, whose views are dis- 
cussed elaborately in the course. There are a great many varieties of 
Kantian idealism. They fill the whole spectrum of opinions between 
positivism and metaphysics. 

The other important scheme of reconciliation is materialism. In its 
original form it was a generalization of mechanistic physics, with no 
metaphysics in it. Later it became obvious that the phenomena of life 
and of human behavior could not be interpreted easily in terms of 
Newton’s mechanics. But a great many people felt the relevance of 
keeping to the mechanistic scheme. In contrast to mechanistic ma- 
terialism, Marx and Engels developed the system of dialectical ma- 
terialism, in which “matter” no longer means the matter of mechanics, 
but a substance which possesses all the qualities that are needed to 
account for the evolution of human life, individual and social. The 
importance ascribed to the principles of this system often went beyond 
their value for the description of observable phenomena, and the pos- 
sibility of any change in these principles was minimized. This attitude 
brought a bit of metaphysics into dialectical materialism. After it had 
become the official philosophy of the Soviet Union, the relevance of die 
principles themselves was bolstered by extrascientific reasons. Dialecti- 
cal materialism is discussed in the course in its application to the 
foundations of science. As in Kantian idealism, we find in dial^tical 


257 



modern science and its philosophy 


materialism all varieties from almost pure positivism to almost pure 
Hegelian metaphysics'. ^ 

These discussions of the philosophic interpretations of science are y 
of great importance to the general education of future scientists. The 
interpretations should not be neglected in the teaching of the phi- 
losophy of science, for they are the link connecting science with the 
humanities. They provide the instrument that is used by religious, 
ethical, and political creeds to muster the support of science. F. S. C. 
Northrop emphasized in his recent book. The Meeting of East and 
West* the great importance of the philosophy of science for any 
ethical or political creed. To understand this connection, we have to 
understand the philosophic interpretations of science and the link of 
metaphysical creeds with religious and political creeds. 

The last part of the course is devoted to the description of these ties. 

It has become almost a commonplace that the communist and other 
left-wing creeds have their philosophic basis in materialism, while the 
right-wing groups look for their foundation mostly to some kind of 
organismic metaphysics, for example, to Thomism. It is, therefore, 
very important that the student have a good training in those philo- 
sophic interpretations which have become the bases of political creeds. 

I The combination of philosophic and political creeds is often referred 
to as “ideology.” The student of science does not need to be ignorant of 
this important field. He can take science and its interpretations as his 
door of entrance. For this reason, in the course in philosophy of science 
much attention should be paid to these philosophic interpretations of 
science that have been the basis of ideologies. The most elaborate of 
these interpretations are Thomism and dialectical materialism. Prom- 
inent cardinals of the Roman Church, as well as prominent political 
leaders like Lenin, who were able to look under the surface of things, 
have taken pains to introduce into the teaching of science an interpreta- 
tion that is favorable to their creed. Hence, the future scientist should 
be taught to take advantage of these ties and get a real insight into 
historical and contemporary ideologies. 

The discussion of these ties helps to make the course in the phl- 

* Macmillan, 1946. 


258 



philosophy of science in the physics curriculum 


losophy of science relevant and attractive to the student. Moreover, 
this way of presenting world problems makes it possible to include a 
discussion of the relations between science and religion— a subject often 
regarded as too delicate for a sincere presentation, and so either omitted 
or presented in a conventional way. On the other hand, I have noticed 
that students are extremely interested in this topic, and ask embar- 
rassing questions if it is treated evasively. If religion is discussed among 
the ideologies, the treatment can become sincere, precise and in- 
offensive. It will become clear that in this question also the best way 
of approach is from science through its philosophic interpretation. 

This is one possible outline of a course on the philosophy of science. 
The choice which I made among a great many possibilities is, of course, 
determined by my own background, scientific and personal. A great 
many paths can lead to the same goal. But all possible courses must, in 
my opinion, be based on logico-empirical and socio-psychological an- 
alysis of science. Only by this method can we steer our ship between 
the two reefs of overspecialization and superficiality. 


259 



CHAPTER 



science teaching and the humanities 

1. Special Field and General Education 

T here is a widespread belief that the rising contempt for toler- 
ance and peace is somehow related to the rising influence of 
scientific thought and the declining influence of ethics, religion 
and art as guides of human actions. This contention is, of course, 
debatable. There is hardly a doubt that the causes of war can be traced 
back quite frequently to religious or quasi-religious creeds and very 
rarely to the doctrines of science. The humanities, including religion 
and ethics, have been for centuries the basis of education and the 
result has been, conservatively speaking, no decline in the ferocity of 
men. The scientists have never had a chance to shape the minds of 
several generations, Therefore, it would be more just to attribute the 
failure of our institutions to educate a peace-loving generation to the 
failure of ethical and religious leaders than to impute the responsibility 
to the scientists, 

I do not think, however, that it makes much sense to discuss the 
share of responsibility. For I agree wholeheartedly with the critics of 
science in the belief that the training of generations of scientists in 
mere science, without making them familiar with the world of human 
behavior, would be harmful to the cause of civilization. Whether we 
like it or not, scientists will participate more and more in the leadership 


260 


4 



science teaching and the humanities 


of society in the future. Also there is hardly a doubt by now that the 
contribution of the scientists to our political life has been more on the 
side of peace and tolerance than have the contributions of the students 
of law or government, or, for that matter, of philosophy proper. 

In order to make this attitude of our leading scientists a habit 
among the rank and file, it is important to imbue the future worker in 
science with an interest in human problems during his training period. 
Since for this purpose it is futile to argue for the supremacy of human- 
istic education over science education, the debate "science versus 
humanities” or vice versa is, of course, without point here. But it is 
also of little avail to compel the student of science to take some courses 
in the departments of humanities. According to the record of all the 
people I know, the mentality of the average science student is such that 
he will not sufficiently appreciate these courses, and therefore will not 
assimilate them well. What we actually need is to bridge the gap be- 
tween science and the humanities which has opened and widened more 
and more during the nineteenth and twentieth centuries. According to 
my opinion, this can be done only by starting from the human values 
which are intrinsic in science itself. The instruction in science must 
emphasize these values and convince the science students that interest 
in humanities is the natural result of a thorough interest in science. 

In this way the science teacher will be giving his support to the 
whole cause of general education as well as to his specialized teaching 
of science. 

Everyone who has ever tried to raise his voice for the cause of gen- 
eral education among the faculty members of a university has been 
running almost regularly against one very definite objection: whatever 
of their time the students have to spend in classes on general education 
they have to subtract from the time they devote to specialized work in 
their own scientific field. As this field is, in any case, so vast that it can- 
not be covered during their stay in college, it would be almost a crime 
to curtail this short and valuable time. This attitude is particularly 
strong among the teachers of science proper. 

I am going to discuss the issue “special field versus general educa- 
tion" mainly from the viewpoint of science students. However, I am 


261 



modern science and its philosophy 


sure that the general picture will be about the same in any other field 
of study, in languages, in history, and so on. 

Even the departments of philosophy have kept to a policy of isola- 
tionism. Instead of working toward a synthesis of human knowledge, 
they have proposed a kind of truce between science and philosophy. 
In my opinion, this gap is greatly responsible for the rift in our gen- 
eral education, or, exactly speaking, the gap between science and phi- 
losophy is the most conspicuous part of the gap between science and 
the humanities— and hence the gap between science and the realm 
of human behavior in general. 

This gap is perhaps nowhere so clear-cut and conspicuous as in 
the domain of physical science. Therefore, the battle for the renewal 
of liberal education will not be won without a willing and intensive 
cooperation of workers in the physical sciences. On the other hand, if 
we want the students of the humanities to go in g’adly for general 
education which requires them to take in quite a few helpings of 
science, we must convince them that by learning science they will also 
advance toward a better understanding of human behavior. 

I am going, first, to describe the harm that the rift between science 
and philosophy has done to both of these fields and to the cause of 
liberal education in general. Second, I am going to make some sugges- 
tions as to how this rift can be repaired by removing the causes 
through which philosophy and science have been estranged. 


2 . Philosophic Interest In Physics 

There is no doubt that the public interest in the physical sciences 
is primarily due to their technical applications: television, radar, the 
atomic bomb. When Copernicus suggested that the motion of the 
celestial bodies be described with respect to the sun rather than with 
respect to the earth, this suggestion was quite irrelevant for any 
technical purpose. Yet the public interest and the heat of the debates 
were certainly greater in this than in the case of any new technical 
device. But we need not go back several centuries for examples, since 
we ourselves have been witnesses of the “relativity boom” which arose 
when Einstein advanced his new theory of motion and light. Although 


262 



science teaching and the humanities 


tbis theory seemed at that time very far from any technical application, 
the public interest was in some cases rather hysterical, and there are 
examples of people who were almost killed in an attempt to get into 
an overcrowded lecture room where Einstein in person tried to put over 
relativity to his audience. There is also no doubt that the philosophic 
and even religious implications of such general physical doctrines ac- 
count for the fact that quite a few clergymen have been eager to make 
use of relativity in their sermons. In order to appreciate this situation 
correctly, we must not forget that Newton, during and after his life- 
time, was a popular topic of parlor conversation and that many books 
popularizing Newton were published, some of them especially designed 
for "the use of the ladies.” 

Nowadays we find, not infrequently, books and magazine articles 
written by clergymen, philosophers, or, for that matter, by scientists, in 
which the theories of modem physics (relativity and quantum theory) 
are recommended for their philosophic or religious benefits. We learn 
from these papers that twentieth-century physics has restored the 
place of mind in the universe, that it has reconciled science with re- 
lig on, and that the tide of materialism characteristic of eighteenth- and 
nineteenth-century science has definitely been broken in the twentieth 
century. As “materialism” has always been connected with some politi- 
cal and social systems, these authors conclude that the new physics 
means also a defeat of all political systems based on materialism, by 
which they mean, according to their personal bias. Communism or, oc- 
casionally, Nazism (racism). 

There is no doubt that the correlation between physics and phi- 
losophy has been largely responsible for the great interest in twentieth- 
century physics of wide sections of the general public. The intelligent 
reader who follows the trend of contemporary thought in books and 
magazines, who listens to popular lectures of scientists, preachers, 
philosophers and global politicians, would often have a greater in- 
terest in the general ideas of twentieth-century physics than an average 
student of physics who specializes, say, in radar. Even after gradua- 
tion, a student of physics usually knows very little about the relation 
between physics and philosophy, let alone between physics and human 


263 



modern science and its philosophy 


behavior. He is generally less trained than the educated layman in 
forming a well-balanced judgment on such problems as are daily dis- 
cussed in magazines and lectures about the influence of modern physics 
on human affairs. If a student in high school or, for that matter, in 
most colleges, asks his physics teacher for information about problems 
of this kind, he will hardly get a satisfactory answer. The information, 
if any, will mostly be perfunctory and evasive. Therefore, the graduates 
in physics will rarely be able to advise the general public on questions 
which this public regards as relevant for their general outlook on man 
and the universe. 

This failure of the learned physicist will not stifle the public interest. 
The thirst for knowledge which is not quenched by the scientists will 
be assuaged by people who are ignorant in science but know how to 
give answers that flatter the wishes of the majority of people. Thus the 
longing for knowledge of large sections of the public will become grist 
for the mills of some organized propaganda groups. 

The textbooks of physics mostly claim to stick to the facts and to 
exclude “idle philosophic talk.” But actually, they formulate the gen- 
eral laws of nature often in such a way that no physical facts whatso- 
ever can be logically derived from these laws. This means that they 
really formulate not physical but purely metaphysical laws. 

Thus, the physical sciences provide very good examples from which 
students can learn that the expression “sticking to the facts only” is 
frequently used as a pretext for avoiding all logical analysis, and there- 
fore for favoring all kinds of obsolete prejudices. What one should 
reasonably mean by "sticking to the facts only” is to make only state- 
ments that can be checked by experience, that is, by observable facts. 
This habit is certainly of great use in debunking empty slogans and 
bigotry in politics or religion. 

As “sticking to the facts” is the slogan of traditional physics teach- 
ing, “ignoring the facts” is a slogan cultivated in the traditional teach- 
ing of mathematics. Both these slogans are logically legitimate within 
a restricted domain of thought. However, on occasion, the students 
have to learn the limitations of these slogans; otherwise, the meaning 
of the most important laws of nature cannot be made clear to them. 


264 



science teaching and the humanities 

and the very goal of general education on the basis of science would 
be frustrated. 

3. Chance Phlloiophles 

Without an understanding of the tie-in between mathematics and 
physics, the student misses the best opportunity of grasping the most 
important trait of human knowledge: the relation between sense ob- 
servation and logical thinking. If this bridge between the fields is not 
built by a thorough analysis of the empirical and logical procedure in 
science, that is, by a systematic philosophy of science, the necessity for 
it is so overwhelming that it will be built anyway. This will be done 
mainly by some obsolete but popular philosophy which will replace 
the thoroughly logical analysis of science. It is noteworthy that, in 
practice, crude empiricism in science, without critical analysis, has 
often made possible the fiourishing of crude metaphysical systems. 

Quite a few great thinkers who belonged to very divergent schools 
of thought have been unanimous on one point; if a scientist believes 
that he has no philosophy and keeps tightly to his special field he will 
really become an adherent of some “chance philosophy,” as A. N. 
Whitehead puts it. This great contemporary metaphysician with a solid 
scientific background assures us that for a scientist deliberately to 
neglect philosophy 

is to assume the correctness of the chance philosophic prejudices imbibed 
from a nurse or a schoolmaster or current modes of expression. 

We find complete agreement with this opinion in a statement of 
Ernst Mach, a philosopher and eminent scientist who was the most 
radical enemy of all kinds of metaphysics. He says, about obsolete 
doctrines of philosophers, that they “have survived, occasionally, much 
longer within science where they did not meet such an attentive criti- 
cism. As a species of animals which has been badly adjusted to the 
struggle of life has survived sometimes, on a remote island where there 
have been no enemies, obsolete philosophy has survived within the 
borders of science.” 

As a third and again very different type of thinker we may quote 


265 



modern science and its philosophy 


Friedrich Engels, the lifelong collaborator of Karl Marx, who was par- 
ticularly interested in the consequences of obsolete philosophy in social 
and political life. He says: 

Natural scientists may adopt whatever attitude they please, they will 
still be under the domination of philosophy. It is only a question whether 
they want to be dominated by a bad fashionable philosophy or by a form 
of theoretical thought which rests on acquaintance with the history of 
thought and its achievements 

One thing seems to be certain: if we try to eliminate from, say, 
physics, all teaching of the philosophy of science, the result will be not 
a crop of scientifically minded physicists, but a flock of believers in 
some fashionable or obsolete chance philosophy. 

Among science students, the students of engineering are those who 
get traditionally the worst training in philosophic analysis. They often 
absorb science, stripped of its logical structure, as a mere collection of 
useful recipes. Is it only a coincidence that the students of engineering 
have on the whole been more impressed by empty political slogans 
(like Fascism) than the students of “pure” science? There is no doubt 
that general slogans play a role in politics s'milar to the role that gen- 
eral principles play in science. If someone is trained to understand 
to what degree general principles like conservation of energy or rela- 
tivity are based on confirmable facts and how far on arbitrariness and 
imagination, he is more immune to the political slogans of demagogues 
than a student who has been trained only to record his immediate ex- 
perience and to regard the general laws as gifts dropped from heaven 
for helping him to bring some order into his record sheet. 

Practically, the separation between science and philosophy can be 
kept up strictly only during a period in which no essential changes in 
the principles of science take place. In a period of revolutionary 
changes the walls of separation break down. In Whitehead’s statement 
quoted above, he makes particularly the point that the lack of phi- 
losophy of physics among the physicists may be harmless in a time of 
stability, but during a period of reformation of ideas this lack will lead 

^F. Engels, Dialectics of Nature, English translation by C. Dutt (New York: 
International Publishers, 1940), p. 243. 



science teaching and the humanities 


unavoidably to the chance philosophy of which we spoke. Our own 
age, with the rise of relativity theory and quantum theory, is an obvious 
example. These new fields have actually become, not only for the lay- 
man but also for the physicist of average training, a kind of mystery. 

Different methods have been used by physics teachers to dodge the 
issue of giving to their students a coherent picture of the laws of nature. 
The simplest thing to do is to stick as closely as possible to the descrip- 
tion of physical devices and the presentation of mathematical computa- 
tions. This way of teaching has given the nonphysicist the impression 
that the science of physics, which has been, historically, the spearhead 
of enlightenment, has become in some cases a source of obscurantism. 
Quite frequently physics has actually been used to attack belief in 
human reason and to bolster belief in irrational sources of truth. This 
misuse had its basis certainly in the failure of many books and instruc- 
tors to give a logically consistent interpretation of the physical meaning 
of the formulas that express the most general laws. 


4, "Profatsional Philoiophy" 

Besides the departments of the special sciences there is in most 
colleges a department of philosophy, which is to counteract the ex- 
treme specialization. It is devoted to the task of investigating the 
foundations that are common to all the special sciences. According to 
our previous argument, the average instruction in the special sciences 
has not achieved the goal of giving to the student an understanding of 
the place of his science in the whole of human knowledge and human 
life. Let us now inquire how the average instruction in philosophy has 
done the job which has been ignored by the instruction in the special 
sciences. As a matter of fact, philosophy (as taught in most depart- 
ments of philosophy) has become a special science itself which is more 
separated from mathematics, physics, or biology, than these branches 
are separated among themselves. The width of the gap that has 
separated science from philosophy became noticeable when the rise of 
completely new theories like the theory of relativity produced a con- 
fused situation among the scientists. The contribution of the philoso-^ 
phers trained in their special field toward a clarification of the new con- 


267 



modern science and its philosophy 


cepts and their integration into the whole system of our knowledge 
has been all but negligible. The students of philosophy trained in the 
traditional way have mostly studied the theory of relativity and quan- 
tum theory from superficial popularizations which were written by 
“physicists” who, in turn, had no training whatsoever in the logical 
analysis and philosophy of science. Therefore, their popular writing is 
imbued with their “chance philosophy” which they have picked up 
somehow. Concepts like space, time, causality are used according to 
these “chance philosophies.” In this way, again, their own traditional 
and sometimes obsolete philosophy has been returned to the philoso- 
phers in the disguise of the gospel of “science.” 

To form an estimate of the width and the depth of the abyss which 
we have mentioned again and again, we have only to make an attempt 
to locate a philosopher who has a “clear and distinct idea” of, say, the 
real issue in the old conflict between Copernicus and the Roman 
Church, let alone of the conflict between the Newtonians and Einstein. 
We would find very few. But it seems ob\'ious that nobody can grasp 
the philosophic meaning of an issue in the history of human thought if 
he does not understand the issue itself— and by “understand,” I mean 
“thoroughly understand.” 

Among philosophers the apology is current that it is just impossible 
for them to have an exact insight into a scientific issue because the 
sciences have become, in our time, so highly specialized that only the 
specialist can have a thorough understanding. But if this is so, how 
again can one have a philosophic judgment about an issue that one 
understands only superficially because the matter is too complicated? 
In this situation a great many philosophers have chosen to establish as 
their redoubt a special field of philosophy outside the field of science. 
To master this field one supposedly needs only an acquaintance with 
the prescientific knowledge that is familiar to the man in the street. 
According to this program of action, the philosopher investigates the 
concepts and beliefs that are the logical basis from which the experi- 
ence of our everyday life can be derived. On this level we make free 
use of words like “time,” “space,” “existence of external objects,” in 
the sense in which the man in the street uses these expressions. The 


268 



science teaching and the humanities 


special sciences like mathematics, physics, biology, as isolated branches 
of knowledge, are taken for granted and the policy of nonintervention 
is upheld. These recognized special sciences have been bom somehow. 
They thrive happily without bothering about philosophic analysis. 
The philosopher wants them to be happy in their innocence and not 
to intmde into his “living space,” which is located between and above 
and below the domain of these isolated special sciences. 

Actually, these autonomous sciences exist only in the oversimplified 
scheme set up by a large group of philosophers. The domain between 
mathematics, physics, biology, history, is exactly of the same stuff and 
has exactly the same logical structure as the domain within physics or 
within mathematics. The borderlines between the special sciences are 
drawn only for the sake of the division of labor and not for any pro- 
found philosophical reasons. The special fields of physics and chemistry 
were regarded for centuries as being of an essentially different nature, 
since physics has to do only with quantitative changes while chemistry 
inquires into qualitative or even substantial changes. Today we have 
between physics and chemistry two new special fields— physical chem- 
istry and chemical physics— which replace the mysterious something 
that was supposed to be the philosophic link between physics and 
chemistry. 

The schools of thought that have advocated the separation of phi- ' 
losophy from science have certainly tended to cooperate in the integra- 
tion of the sciences, but they perform this job by using as binding ma- 
terial some prescientific stuff, while we leam from our last example 
that the binding material between the special sciences is itself a full- 
fledged science. But another school of thought, which claims to be very 
up-to-date, takes an attitude that we may call an attitude of defeatism. 

It leaves the special sciences untouched and autonomous. But according 
to this school, philosophy does not even attempt to fill the gaps between 
these special sciences but plans to build up a completely separated 
stratum of knowledge “beyond science.” This "knowledge” is claimed to 
be completely independent of any advance of science proper, for it is 
based only on the prescientific experience of mankind. 

We may distinguish two groups within this school. Both insist thaf 



modern science and its philosophy 


when the scientist has done his job as thoroughly as he can, the phi- 
losopher’s job begins. When, for example, the physical laws of motion 
are established by the scientist, the philosopher, says the first group, 
steps in and puts his particular questions. The scientist has formulated 
by his laws how motion takes place, what it is like, and so on. But the 
philosopher wants to know what motion is, with the emphasis on the 
“is.” While the scientist explores the observable attribute of motion, 
the philosopher wants to find out the “being,” the “essence,” of motion. 
This essence of motion can be discovered on the basis of our pre- 
scientific knowledge about motion and cannot be affected by any 
further advance in our science of mechanics. To this group belongs the 
present-day neo-Thomist. 

The second group starts also from the special sciences as having ac- 
complished their job. But instead of looking for the “being” of things 
this group claims that these special sciences take some “presupposi- 
tions” for granted without investigation, such as the existence of ma- 
terial bodies, the law of causality, the law of induction. Then, they say, 
the philosopher has to step in and investigate whether these presupposi- 
tions are correct. When I hear this claim, I have sometimes the feeling 
that the shoe may be on the other foot. For quite frequently scientists 
investigate the presuppositions that philosophers have taken for 
granted without investigation. The founders of non-Euclidean ge- 
ometry, Gauss, Lobatchevski, and Bolyai, doubted the axioms of 
Euclidean geometry. Einstein doubted the axioms of Newtonian 
mechanics, while a great many philosophers believed in these axioms 
as eternal truths. Moreover, it is quite debatable whether “presupposi- 
tions” like the existence of material bodies really play any role in 
science and whether presuppositions that do play a substantial role 
can be investigated by any method which is not scientific itself. What- 
ever may be our final judgment about this investigation of presupposi- 
tions, the practical effect of this philosophic school is again the es- 
tablishment of philosophy as a special science besides mathematics, 
physics, economics— and the perpetuation of a wide gap between 
science and the humanities in our educational system. 

The role of philosophy as an integration of human knowledge is 



science teaching and the humanities 


ignored, or, at least, neglected; consequently, the educational values 
intrinsic in mathematics or physics are not exploited. These special 
sciences are reduced to the status of useful knowledge without truth 
value while, on the other hand, “philosophy” becomes a type of dis- 
course without contact with the advance of science and, therefore, 
without contact with the evolution of human intellect. 

From these considerations, it seems obvious that the traditional 
teaching of philosophy may have contributed considerably toward 
sharpening the thinking of students and giving them a certain touch of 
sophistication, but has certainly made little contribution toward the 
synthesis of human knowledge which should be the chief goal of 
liberal education. 


5. Neo-Thomiam and Dialectical Motertolism 

There is a suggestion which has been widely discussed during 
recent years— the idea of Robert M. Hutchins, Chancellor of the Uni- 
versity of Chicago. The essential point of his thesis is that we have to 
base the integration of knowledge taught in our colleges on the last 
available synthesis in the history of thought, on a kind of “standard 
tradition.” According to this group, the spokesman of which has been 
the philosopher Mortimer ] . Adler, the last system in history that has 
really achieved a synthesis of science, ethics, politics, and religion is 
the philosophy of St Thomas Aquinas. His Summa Theologiae 
and his Summa CathoUcae Fidei contra Gentiles present a coherent 
system in which, from the same set of principles, not only astronomy, 
psychology, ethics, and politics are derived logically, but also the be- 
havior of the angels— for example, whether the speed of their flight is 
finite or infinite. 

It seems, of course, debatable whether actually Thomism is the last 
coherent system that has attempted or achieved such a sweeping 
synthesis. Some people would, certainly, claim that the philosophy of 
dialectical materialism, which is the ofBcial basis of education in the 
Soviet Union, is ajlso a set of principles from which are derived not 
only physical science but also the laws of history and sociology. Just as 
well as Thomism this more recent system claims to give guidance not 


271 



modern science and its philosophy 

only in scientific research, but also in the question of what is a “good 
life.” 

The basic contention of Hutchins and his group is that a synthesis 
which may not be perfect is preferable to no coherent synthesis at all. 
There is no doubt that it is the chief asset of Thomism that such 
disparate subjects as astronomy and theology can be regarded as con- 
clusions from one and the same set of principles. But disregarding 
theology, hardly anyone would claim that Thomism is a good system 
from which to derive an answer to the question whether the New- 
tonian or the Einsteinian mechanics is preferable. 

In the same way, the chief asset of dialectical materialism is the fact 
that the laws of physics are derived from the same principles as the 
laws of human societies. We learn from the textbooks of dialectical 
materialism that, for instance, the law of the transition of a capitalistic 
society into a communistic one follows from the same principle as the 
transition of water into steam. Both are conclusions drawn from the 
dialectical principle that quantitative changes eventually become quali- 
tative changes. But if we are not interested in the synthesis of physics 
and sociology into one set of principles, hardly anyone would claim 
that dialectical materialism is the best foundation of physical science— 
j-for example, the most helpful interpretation of the evaporation of 
liquids. 

Dialectical materialism has, as a matter of fact, nowhere been 
chosen as a basis of education except in countries where the govern- 
ment has been committed to Marxist economic and political principles. 
In this case, there is clear advantage in having these principles linked 
up with the laws of physical science by a common set of principles. 
With the same right we can assume that Thomism is not commendable 
as a basis of education except where the government is committed to 
the political and religious doctrine of the Catholic Church. For it will 
enlist science by regarding science, politics and religion as derived from 
common principles. 

There is, on the other hand, no doubt that in an education which 
emphasizes the integration of human knowledge, much more attention 
than usual should be given to the systems that historically have per- 


272 



science teaching and the humanities 


formed such an integration, however we may judge the actual political 
and religious way of life which is coherent with this system. The student 
should get a good and unbiased presentation of both Thomism and 
dialectical materialism as syntheses of human knowledge. But to make 
either of these systems the main or exclusive basis of education in the 
philosophy of science can be justified only if a particular political and 
religious indoctrination is intended. 


6. Integration of Science and Philoiophy 

Before we can set up a constructive plan for bridging the gap be- 
tween science and philosophy and, as a result, between science and the 
humanities, we have to remove the chief obstacles blocking the way 
toward this goal. As we have learned, the two principal obstacles are, 
first, the exaggerated belief of scientists in specialization which some- 
times leads even to a prejudice against general ideas and, second, the 
recent tendency of the schools of philosophy to establish “philosophy” 
as a new special science, instead of working on the synthesis of 
knowledge. 

The negative attitude of many scientists is based on their convic- 
tion that any trespassing beyond the limits of one’s own field would 
lead to unavoidable superficiality. Therefore, the genuine scientist has 
to mind his own business and keep within the fences of his own de- 
partment. There is, of course, a grain of truth in this argument of 
avoiding superficiality. However, it does give only one side of the 
picture, for the advance of science has revealed not only more and more 
complexities in science, but also more and more cross-connections be- 
tween the “isolated” special branches. By this fact it has become much 
easier than formerly for one man to grasp the findings of several spe- 
cial fields. We have only to consider die example of physics and 
chemistry. 

If we want to get a sound judgment of how, despite the abundance 
of factual material, to acquire a thorough knowledge across depart- 
mental lines, we have to ask, for example, how some people have man- 
aged to become experts in a field like biophysics. They certainly did 
not do it by a thorough study of the whole of physics plus the whole 


273 


O 



modern science and its philosophy 


of biology, for this cannot be achieved in one lifetime. Instead, they 
acquired a balanced survey knowledge in both fields, physics and 
biology, and tried to acquire a really thorough knowledge in those 
parts of physics and biology which are relevant for the interaction of 
the phenomena of life with physical phenomena. As a matter of fact, 
the behavior of the scientists who have worked within a traditional 
field like physics has not been different. An average physicist will sur- 
vey first general physics and then obtain a detailed acquaintance with 
his special field within physics. If he wants to become a biophysicist 
his survey information has to be broader, but his field of special in- 
terest need not be larger than the special field of an ordinary physicist. 
Moreover, to be quite truthful, the average physicist learns some part 
of physics outside his special field only through popular generaliza- 
tions. This is frequently true for the theory of relativity. The individual 
physicist is, of course, not to blame for this situation, for without using 
popularizations he would not be able to get any information about 
important fields of his science. 

From these remarks it becomes obvious what must be the train- 
ing of the “philosopher of science” if his goal is a synthesis of human 
knowledge. He has to acquire a survey knowledge of several sciences 
and a thorough and precise knowledge of those parts of each special 
field that are relevant for the relations across the borderlines and for 
the relation between science and human behavior. 

Some people may object that a survey knowledge would not be 
sufficient, for one cannot know what part of science will be relevant 
for the purpose of philosophy before the integration has been actually 
achieved. There may be some truth in this argument, but it proves too 
much, For according to this argument, every physicist must have a 
thorough knowledge of the whole of physics; otherwise he cannot know 
what knowledge may be relevant for his special field of physics. Noth- 
ing can be done about it and he just has to take the risk in his training 
as a physicist. He will learn by and by to smell what is relevant and 
what is not. No greater effort is, in principle, required of the philos- 
opher who wants to acquire a training in the philosophy of science. 
There is no doubt, however, that evenfa. survey knowledge in the sci- 


274 



science teaching and the humanities 


ences^ill take so much of his time that he will not be able to get the 
training that a philosopher has to get if he goes into “philosophy as a 
special science.” 

But it may be sufBcient for a student who specializes in the philos- 
ophy of science and wants to take his Ph.D. in philosophy to get along 
with a survey knowledge in the history of philosophy, without learn- 
ing the details of all the opinions that have been uttered through two 
or three thousand years. Every philosopher of science should, of course, 
be familiar with the ideas of the great thinkers like Plato, Aristotle, 
Thomas Aquinas, Leibnitz, Descartes, Kant, Nietzsche. But it is per- 
haps sufficient if this special candidate becomes familiar with the 
language of these men and knows how to locate their ideas within the 
great stream of the evolution of scientific thought. This would leave him 
time and, more important, the leisure to acquire a good survey of the 
physical and biological sciences. He would concentrate his effort on 
those parts of these sciences which are the most relevant for judging 
the borderline problems arising between the special sciences and be- 
tween science and traditional philosophy. He would concentrate, for 
instance, in mathematics on problems like the “truth of non-Euclidean 
geometry”; in mechanics on the role of “absolute motion”; and, in par- 
ticular, on the ties between mathematics and physics, such as the dis- 
tinction between mathematical and physical truth of geometric axioms. 
He would, of course, try to acquire a thorough understanding of Ein- 
stein’s theory of relativity, of Heisenberg’s principle of uncertainty, of 
Bohr’s complementary concept of nature, and so on. In traditional 
philosophy he would try to understand the approaches of different 
schools to the question of what is the precise borderline between 
physics and philosophy. He would try to learn the answers of the great 
philosophers to questions like: What is the logical status of the general 
laws of nature? Are they a result of experience or of reason or of what? 
What are the roles of chance and of causality in the general laws of 
nature and in their application to observed phenomena? 

Teachers of philosophy with a similar type of training could give to 
the students reliable information about the problems of the “philos-^ 
ophy of science” and of the “integration of sciences.” 


275 



modern science and its philosophy 


But tihen we are confronted by a further task. If we know even the 
problems, do we know also the solutions? What should we present to 
the student as the result of the integration of science? One should give 
him reliable guidance without providing him with a “chance philos- 
ophy” which may be either the result of an old and now obsolete tra- 
dition or just the fashion of a year and a certain social group. 

There is no doubt that the integration of knowledge on the college 
level can be promoted among the students only by the use of philosophi- 
cal and historical argument. However, the starting point has to be 
living science itself. Philosophical and historical discourse must ema- 
nate from this source. There are quite a few good reasons for this, but 
it may be sufficient to consider the practical reason that in no other way 
can philosophy and history be made palatable to the student of science, 
and he will fail to appreciate this unusual food if he has no appetite 
for philosophical and historical ideas. It would be, of course, a poor 
teaching method just to add to the traditional presentation of science 
some philosophic spice or sauce. We have rather to give to the presenta- 
tion of science itself a philosophic touch. 

The teacher of the special sciences will perhaps be afraid lest time 
would be wasted by such a treatment. The student would pay for this 
philosophical and historical touch by a deficiency of information in 
science proper. But it seems to me that this new approach will rather 
save time. For by this method a great many laws of physics, for ex- 
ample, could be much more attractively presented to the students than 
by traditional methods. However, I do not mean that the approach 
should be made by one of the numerous metaphysical systems that 
have been invented during the ages for the purpose of an integra- 
tion of human knowledge. Every attempt of this kind would introduce 
very questionable doctrines into the teaching of science and would 
lead to disaster. We have to make use of the philosophic argument 
that has grown up on the soil of science and has been fed with the 
blood of science. We must never forget that metaphysics divides people 
and science unites them. 

If we try to build the bridge between science and philosophy, our 
first step will be to present to the students their own special science as 
a chapter in the book of human knowledge. Every scientist is con- 


276 



science teaching and the humanities 


fronted with the amazing fact that it is possible to derive from a few 
simple principles by means of logical argument a wide range of facts 
which can be checked by actual observations. The existence of these 
principles allows us to put the phenomena of nature into our service, 
for they enable us to construct methods by which the outcomes of physi- 
cal processes can be predicted from the start. 

Philosophy of science is concerned with the nature of this method 
or device which man has invented in order to bring about the predic- 
tion of physical phenomena. To have a certain understanding of this 
device is a basic requirement for everyone who wants to understand 
the history and the behavior of mankind in past centuries, and in our 
own. 

An understanding of the logical structure of science is a long step 
toward the understanding of the meaning of statements in any domain 
outside science proper and, indeed, toward judging truth of any kind. 
In fields like ethics, politics or religion, we have also to distinguish 
clearly between the factual content of a certain doctrine and the sym- 
bolic language in which the statement of this doctrine is couched. The 
example of physical science is a guide in a more difficult world and 
will help us to disentangle statements of religious or political principles 
with respect to whether they are really statements about observable 
phenomena or only attempts to use a certain type of symbol. 

In physics this analysis is comparatively simple and not so loaded 
with emotional and egotistic elements. If someone asks people in the 
strongest language to “follow the voice of their conscience” or to “fol- 
low the will of God," this bid will be empty if he is not able to describe 
the criteria by which we can know whether a “voice” is actually the 
voice of our conscience or how actually to find out the will of Cod. The 
student of science who has been trained in the "understanding” of 
science will immediately turn his attention not so much to the strength 
of the language, but to the question of who is authorized to interpret 
the will of God. 


7. Role of the Human Mind 

By logical and empirical analysis the student will learn that the 
principles of science are neither “proved by reason” nor “inferred by 


277 



modern science and its philosophy 


induction from sense observation.” They are a structure of symbols 
accompanied by operational definitions. This structure is a product 
of the creative ability of the human mind and consists of symbols 
which are products of our imagination. But the truth of this struc- 
ture can be checked by observations that can be described in every- 
day language. By logico-empirical analysis the creativity of the hu- 
man mind emerges as the primary factor in science. Thus the stu- 
dent will learn that the role of this creativity in science is by no 
means inferior to its role in the humanities and even in art or religion. 
And we now can understand that the emphasis on science teaching will 
no longer interfere with interest in the humanities but will rather sup- 
port it. 

I However, the role ascribed to the human mind by logico-empirical 
analysis does not exhaust the contribution of science teaching to the 
understanding of the human aspect in our picture of the world. For 
by logico-empirical analysis the role of the human mind is only hinted 
at in a rather abstract way. But our imagination and inventiveness are 
much too limited to enumerate and discuss all possible principles that 
the creative ability of the scientists may set up in order to derive the 
wide range of phenomena of our experience. For this we have to study 
the principles that have been actually set up in history. We have to 
complement our logico-empirical analysis— where “empirical” implies 
individual experiences— by “historical analysis,” which is empirical not 
for the individual but for the human race. The history of science is the 
workshop of the philosophy of science. We have to teach the student 
all the relevant principles that have been set up in the course of history. 
And we mean by history extension in time as well as in space, the de- 
velopment of structures of science through the ages and over the sur- 
face of our globe. 

In this way the logico-empirical analysis gains life and color and 
becomes a living link between science and the evolution of the human 
race. The average textbook of physics tells us very little about the evo- 
lution of the principles of this science, except some dates of anniver- 
saries. Very often these books speak about ancient and medieval sci- 
ence in a derogatory way; they claim not to understand why for ages 


f 


c> 


278 



science teaching and the humanities 


people were not able to discover such a simple law as the law of in- 
ertia, which today every schoolboy knows is an obvious result of our 
everyday experience or is even self-evident. But despite these smug 
remarks the same textbooks are not able to formulate this law of inertia 
in a satisfactory way. They even block the understanding of this and 
similar principles. For it is clear that a principle which intelligent men 
have not found through centuries cannot be as obvious as the state- 
ment presented by these books as the law of inertia. This complacent 
attitude imperils even the understanding of the evolution of thought 
and helps to spread the spirit of intolerance and bigotry among the 
students, while an attitude of adequate logico-historic analysis would 
contribute toward good will between people of different backgrounds 
and different creeds. 

The best way to help the student to understand the steps in the 
evolution of human thought is to present to him in elaborate detail the 
chief turning points in the evolution of science, with the emphasis not 
so much on the discovery of new facts as on the evolution of new prin- 
ciples of change in the symbolic structure. It would be, for example, 
of the greatest importance to discuss thoroughly the conflict between 
Copernicus and the Homan Church (or, for that matter, the Lutheran 
Church), I think that every student of science and the humanities 
should have a clear understanding of this issue, which was one of the 
greatest and most interesting in history. If this subject were discussed 
thoroughly and competently, the student could get a good understand- 
ing of the eternal conflict between established patterns of presenting 
the facts and attempts radically to alter the symbolic structure of 
science. He would learn that the tendency to preserve the old pattern 
of presentation is often disguised under the name of “common sense,” 
and how the appeal to common sense has been used in the history of 
mankind to cloak the interest of established governments and churches. 
For, as he would learn in particular, the role that the interaction among 
science, philosophy and religion has played in the justification of po- 
litical aims is very great. 

Equally, students of science and philosophy should learn exactly 
what were the issues between Descartes and Newton and between- 


279 



modern science and its philosophy 


Newton and Leibnitz. From these disputes has arisen what we now 
call the classical physics of the nineteenth century, which until today 
has been the basis of the training in science that our students get in 
colleges of engineering or liberal arts. To grasp these issues would help 
them to understand our present science as a dynamic living being. This 
would not happen if they were confronted only with the desiccated 
and artificially stuffed skin of science that is presented in most of the 
current textbooks. 


8. Science and Political Ideologies 

If the students get an understanding of the earlier turning points in 
science, it will be much easier for them to grasp exactly the meaning 
of the turning point around 1900, when our twentieth-century science 
was bom. 

This last turn has been dramatized by the phrases “crisis of classical 
physics” or “decline of mechanistic physics” or “refutation of material- 
ism.” If one has been trained to analyze the nature of a “turning point 
in the history of science,” one will be less inclined to believe that the 
“crisis of classical physics” is a “crisis of rational thinking” or even a 
justification of an irrational approach to science. 

As we have already mentioned, it is not suflBcient to approach these 
turning points of human thought by logico-empirical analysis only, for 
the human mind is not strong enough to carry out an exact analysis of 
such a complex stmcture. One has to study classical physics as an ex- 
.tinct organism, which grew up against immense obstacles, defeated its 
opponents, and then turned out to be no longer fit for survival. With 
this training one would have a clear understanding of, say, the broad 
analogy between the fight of medieval philosophy against Copernicus 
and the fight of modern Newtonian philosophy against Einstein. 

Students who have this kind of logical and historical training will 
easily see through attempts to exploit the “breakdown of Newtonian 
physics” and the “defeat of materialism” in order to justify a return 
to ancient “organismic science.” They will be, moreover, on their 
guard against attempts to exploit this “crisis of thought” in a fight 
against liberalism and democracy, or, for that matter, against all pro- 


280 





science teaching and the humanities 


gressive trends which have been historically labelled "materialistic’’ or 
“atomistic’’ or “mechanistic." 

By this approach the student of science would be led in a natural 
way to an understanding of the struggle among rival ideologies. It will 
be a great attraction for him to approach these problems starting from 
the role that has been played by his own special field. The student of 
science will get the habit of looking at social and religious problems 
from the interior of his own field and entering the domain of the hu- 
manities by a wide-open door and not by the rear door of some isolated 
humanity course which he may take for "distribution” He wiU need 
neither a spoon feeding of trivial information nor a stuffing with techni- 
cal material that is of no real profit for his general education. 

There is no better way to understand the philosophic basis of politi- 
cal and religious creeds than by their connection with science. The 
student who understands the relation of his science to these creeds 
has an access from an inside track. He will easily and confidently cross 
the bridge between science and the humanities. 

The attentive student of science will notice soon that the tradi- 
tional symbols of science have a life of their own. They persist in a 
changing world where the scope of science is continually growing. 
This point is made particularly clear by focusing the attention of the 
students on the turning points in the evolution of scientific thought. 

The student will learn, for example, in what sense materialism has 
been encouraged by the physics of the nineteenth century and how 
this in turn was anticipated to a certain degree by the Epicurean School 
in old Greece. He will learn how the transition from medieval physics 
—which was based, in its turn, on the Aristotelian school of Greece— 
to the physics of Galileo and Newton found its continuation in the 
school of Laplace at the end of the eighteenth century at the time of 
the great French Revolution. He will appreciate, then, how the fight of 
Newtonian (mechanistic) physics against Aristotelian (organismic) 
physics became connected with the fight of liberalism and tolerance 
against feudalism and bigotry. 

He would thus understand that what scientific and corresponding 
political (ideologic) issues have in common is the use of the same sym* 


281 



modern science and its philosophy 


bols, with their wide range of connotations. In this way the student of 
science would learn to appreciate the great value of symbols in the 
history of human thought and, for that matter, in the history of human 
behavior. 

Whoever has understood these historic issues correctly will attain a 
sound judgment regarding the last great transition, around 1900, when 
mechanistic physics had to give way to a more general approach. The 
transition from the nineteenth-century to the twentieth-century physics 
culminated in the relativity and quantum theories which, in turn, have 
led to new philosophic slogans describing this transition as an “over- 
throw of the concept of absolute time and space” and an “overthrow 
of physical determinism.” The student who has been through the train- 
ing in logico-empirical and historical analysis will assess the attempts 
that have been made to exploit the new physical theories for the benefit 
of particular religious and political ideologies. He will see through 
the argument by which the “overthrow of eighteenth- and nineteenth- 
century deterministic physics” has been used in the fight against liber- 
ahsm and tolerance, since these creeds had grown up in a period of 
mechanistic and deterministic science. He will understand that the 
breakdown of mechanistic physics did not actually imply a return to 
organismic physics, which was, historically, connected with the political 
and religious doctrines of the Middle Ages. He will understand why 
twentieth-century Fascism has gladly interpreted the “crisis of physics” 
as a return to organismic physics which could provide a “scientific” 
support for a return to some political ideas of feudalism. 

But, above all, the well-trained student will understand the para- 
mount fact that, actually, mechanistic physics was replaced^ not by 
any organismic physics, but by an entirely new approach to science by 
logico-empirical analysis, which has been in the twentieth century the 
starting point of all the new physical theories. 

If science is taught in this way, the emphasis on science and tech- 
nology toiU no longer be an obstacle to a liberal education of the stu- 
dent. The deplorable gap between science and the humanities will not 
arise, let alone widen. On the other hand, the intensive study of science 
as a living being will give to the student of it a profound understand- 


282 



science teaching and the humanities 


ing of the role of the human mind in human action, which is the very 
goal of instruction in the humanities. 


9. Science and the Historical Systems of Philosophy 

By emphasizing the historical evolution of scientific thought the 
student will learn, moreover, that the human mind has not always been 
satisfied with the logico-empirical analysis of science, since this presen- 
tation of science is only satisfactory for the “purely scientific” purpose 
of predicting and mastering the observable phenomena of nature. But 
the phenomena are derived from principles that are couched in sym- 
bols and, as we have already hinted, these symbols have their own 
life, which is to a certain degree independent of the evolution of sci- 
ence proper. These symbols which are created hy the scientists may 
even become occasionally a Frankenstein’s monster. However, as these 
symbols are not unambiguously determined by the scientifically ob- 
served facts, they are strongly influenced by extrascientific factors. The 
choice of symbols is, as a matter of fact, very dependent on the impact 
of current social and religious movements. These influences are largely 
responsible for the decision whether one prefers rigid pieces of mat- 
ter as fundamental symbols (materialism) or whether one builds up all 
concepts from mental elements (idealism), whether one picks as the 
ultimate building stone a nondescript reality (realism), or whether 
one starts from elements that cooperate toward a certain goal (organ- 
icism). Every satisfactory instruction in the philosophy of science has 
to discuss these choices of symbols on the basis of logical and historical 
analysis. The influence of political and religious trends on the choice of 
these symbols should by no means be minimized, as is often done in the 
presentation of the philosophy of science. On the other hand, if “meta- 
physical integrations of science” are discussed, particular attention 
should be given to those integrations that have played a role as bases of 
ideologies. For this reason, doctrines like Thomism or dialectical mate- 
rialism should be carefully and correctly presented to the student and 
more time should be devoted to them than is devoted to some sophisti- 
cated systems that have played only a small role in human life and 
human actions. ‘ 


283 



modern science and its philosophy 


If the foregoing plan is followed, we shall have no more graduates 
in science who have no clear idea of the teaching of men like Aristotle, 
St. Thomas, and William of Occam or, for that matter, of Hegel, Marx, 
and Lenin. The type of science graduate who is without humanistic 
training will disappear just like those who have not even a clear picture 
of what the contribution of Copernicus was to our world. 

The educational value of this type of instruction for science students 
seems to me beyond doubt. However, there is still the question of where 
to find a place for it in the curriculum. The most natural plan probably 
would be to teach the science courses of broader scope according to this 
method. This would hold, for example, for the elementary courses in 
coUege physics, chemistry and biology. Such a start would certainly be 
very stimulating and helpful for the beginning students. However, 
since these students have not the background necessary for the study 
of subtle problems, these "survey courses for beginning students” 
should be complemented by “survey courses for advanced students.” 
These would be appropriately given just before graduation. They 
should answer the questions that were prompted by the elementary 
courses and treat them on a higher level. These new courses should not 
be “superficial surveys” as this tenn is often understood, but should 
give a bird’s-eye view of the results of science, with emphasis on spe- 
cial unsolved problems. These courses could be given according to the 
suggestions of this paper. 

If there are not a suflScient number of science teachers in a college 
who are interested and trained in this plan of instruction, one or two 
“special” courses outside the usual science curriculum should be estab- 
lished, to be given by the few available teachers who have the necessary 
training and inspiration for this task. One may give these courses un- 
der the tide of “philosophy of science” or “foundations of science” or 
“science and the humanities.” 

The present trend toward general education has, in some colleges, 
led to the establishment of science courses for nonscience students. 
The program of these courses emphasizes the bridge between science 
and philosophy or science and the humanities somewhat along the lines 
tliscussed in this paper. In these plans, however, only the nonscience 


284 



science teaching and the humanities 


student will be presented with the educational value of science, while 
the concentrator in science will not be able to give information about 
the role of science in human society to his future pupils or to his 
community in general. The questions regarding science that are most 
interesting to the general public should be answered by a compe- 
tent and responsible man; and this obviously can be only the science 
teacher in the high school or college. 


285 



CHAPTER 



the place of logic and metaphysics in the advancement 
of modern science 


O NE of the most brilliant writers on intellectual history, Carl 
Becker, claims that the most important event in this field in 
modem times was the shift in the place of logic in science. 
According to Becker, the high esteem in which logic had been held by 
the scientist in the time of St. Thomas Aquinas and through all the 
Middle Ages declined in the period of Galileo and Newton. But at that 
time this decline was not yet fully understood. “The marriage of fact 
and reason,” as Becker puts it, “proved to be somewhat irksome in the 
nineteenth century and was altogether dissolved in the twentieth cen- 
tury.” The modem, twentieth-century, physicist lives in an "atmosphere 
which is so saturated with the actual that we can easily do with a 
minimum of theoretical. . . . We have long since learned not to bother 
much with reason and logic.” To describe the spirit of twentieth- 
century physics, which emphasizes facts and minimizes reason, Becker 
says; 

Experiments seem to show that an electron may, for reasons best known 
to itself, be moving in two orbits at the same time. To this point Galileo’s 
common-sense method of noting the behavior of things, of sticking close to 
the observable facts, has brought us. It has at least presented us with a fact 
'that common sense repudiates. 


286 



logic and metaphysics in modern science 


I do not think that Galileo was actually less concerned with logical 
consistency than Aristotle, nor do I think that Einstein’s theory of gravi- 
tation is more factual and less rational than Newton’s. I shall not enter 
here into historical arguments; I shall restrict myself to investigating 
the nature and the bearing of that “most important event in the intel- 
lectual history of modem times” which Becker stresses so strongly. 

The pre-Galilean period, say until about 1600 a.d., adhered to a kind 
of organismic world view founded largely upon the philosophy of Aris- 
totle and the medieval schoolmen. The "mechanistic” conception of 
nature began with Galileo and Newton. Its peak, however, was reached 
in the first half of the nineteenth century. According to Becker, this 
whole period including the Middle Ages, the Renaissance, and the 
eighteenth-century enlightenment, was characterized by the belief that 
there is a rational picture of the world, there is a way by which man 
can comprehend nature by reason. But in the twentieth century this 
belief faded more and more. An acute observer would have already 
noticed this fading in the Galilean period when the confidence in the 
scholastic type of philosophy declined. More and more, the "cli m ate 
of opinion,” as Becker calls it, became less logical and rational and 
more factual and in some ways irrational. 

We shall probably not seriously believe that science has ceased to 
be logical. If we use this word in its technical sense, everyone will 
agree that a science that repudiates logic has never existed and can 
never exist. Now, what did Becker really mean by his “decline of 
reason and logic” in the twentieth century and even in the approaches 
to this century? Why did he believe that the organismic medieval 
world picture was rational as well as the Newtonian and Laplacian 
mechanistic picture, while the twentieth-century picture of physics, 
characterized by relativity and quantum theory, is no longer concerned 
with reason? 

We shall understand this point better if we direct our attention to 
the fact that, according to Becker, this decline of the rational world 
view had already started in the Galilean period and is, therefore, some- 
how connected with the d^line of scholastic philosophy. 'The common 
feature of medieval and of mechanistic physics seems to me to be that 


287 



modern science and its philosophy 


their principles seemed to have a certain plausibility by themselves. 
Medieval science derived all observable phenomena from the principle 
that they are somehow analogous to the well-known phenomena in a 
living organism. Seventeenth- and eighteenth-century science, in turn, 
preferred the analogy to simple mechanisms which are familiar to us 
from our everyday life experiences. Not only were the principles con- 
firmed by demonstrating that the conclusions drawn from them were in 
agreement with observed facts, but also the principles themselves used 
to be directly confirmed by a kind of short cut. If a law of physics was 
in agreement with the organismic or mechanistic analogy, this agree- 
ment accounted for a certain degree of confirmation. A more careful 
analysis would probably lead to the result that the organismic as weU 
as the mechanistic principles of science drew their plausibility from the 
fact that they seemed to reflect faithful pictures of our everyday ex- 
perience. In scientific theories a long chain of intellectual and experi- 
mental work connects the principles of a theory with the protocols of 
observation. The organismic or mechanistic principles, however, could, 
it was believed, be confirmed directly by a very obvious type of ex- 
perience. Everybody knew that a piece of earth tliat is dropped falls to 
the ground. It therefore seemed obvious that the earth as a whole could 
not remain suspended in space and circulate around the sun. It was to 
fall to the center of the universe like a small piece of earth. This tend- 
ency toward the center as the natural place of the heavy element earth 
seems to be a principle of physics with firm roots in everyday experi- 
ence. In the same way, a mechanistic principle like the law of inertia 
seemed to be directly confirmed by the most familiar facts of our daily 
experience. 

In this sense we can say that the organismic as well as the mechanis- 
tic physics were based on principles that could be interpreted directly 
in terms of everyday experience. The validity of the principles did not 
need any further confirmation by the development of more refined 
methods of observation and theoretical argument. Science, in those 
days, was based on principles to which an eternal validity was as- 
cribed. They appeared to be “rational” or “reasonable” or, expressed 
in a loose way, “logical.” The philosophers of the eighteenth-century 


288 



logic and metaphysics in modern science 

enlightenment still felt that theirs was a world view, provided com- \ 
pletely hy reason. The age of enlightenment shared this belief in 
reason with the “dark” Middle Ages. This is the historical point that 
Carl Becker wanted to make in his book. In the nineteenth century, 
according to Becker, a drive was initiated to tell reason that it had to 
know its place, which was considerably lower than previously; but in ' 
the twentieth century, science emancipated itself from reason alto- 
gether. 

After these few remarks about the place of reason in the science of 
past centuries, it is obvious what the characteristic of twentieth- 
century science is. With the rise of non-Euclidean geometry, the 
physics of relativity and the mechanics based upon the de Broglie 
waves, the basic principles of physical science were no longer a direct 
formulation of our everyday experience; they were no longer obvious 
and plausible to common sense. Their only justification consisted now 
in their property of yielding observable facts by means of a chain of 
logical conclusions. The illusion disappeared that the principles of 
science were of eternal validity or could at least be interpreted as con- 
clusions from such principles. Therefore, one ought not to say that 
twentieth-century science has no use for logic but rather that it has no 
use for metaphysics. To interpret the principles of science as results of 
our common sense leads to the opinion that they are self-evident and 
cannot be refuted by further empirical checking. This belief is the very 
core of the metaphysical interpretation of science. 

One should rather reformulate Becker’s argument as follows: In the 
period of Galileo and Newton the firm belief in the metaphysical 
foundation of science faded a little. Mechanistic physics was less im- 
bued with metaphysical argument than was medieval, organismic, 
science. But the belief in Newton’s laws as results of the simple ex- 
perience of everyday life had been bolstered up during the eighteenth 
and a great part of the nineteenth centuries, although it was by no 
means a common opinion among Newton’s contemporaries. With the 
new physical theories of the twentieth century— non-Euclidean geome- 
try, relativity and the quantum theory— the belief practically disap- 
peared that the basic principles of physics ought to be plausible ac- 


289 



modern science and its philosophy 


cording to the criteria of common sense. The metaphysical conception 
of science lost ground and the logico-empirical interpretation of the 
scientific method became the method that was actually used. We can 
see in the evolution of science from the seventeenth century to the 
twentieth century the gradual decline of metaphysics in favor of a 
positive conception, if we want to use the terminology of August 
Comte. 

I am using the term metaphysics here with a positive and precise 
meaning; direct interpretation of the basic principles of science in 
terms of common sense or everyday experience. I think that it is not 
sufficient to characterize metaphysical statements as “meaningless.” 
There are a lot of meaningless statements that are not at all “meta- 
physical.” By using the term in the way I suggested we can probably 
cover all the statements which have been made with the claim to be 
metaphysical. Metaphysics, according to our way ot speaking, is cer- 
tainly meaningless from the scientific viewpoint because the terms 
“true” or “false” cannot be applied to these statements. Charles Morris 
speaks, however, of a “metaphysical discourse.” I agree with him in 
the sense that I regard metaphysics as a direct interpretation of 
scientific principles in terms of the language of everyday life experi- 
ence. “Interpretation” means translation. Metaphysics attempts a trans- 
lation of the basic principles of science, but not according to a strictly 
fixed dictionary; the univocal relation between a term and its transla- 
tion has been replaced by an analogical relation. But we cannot tell by 
any exact criterion what is a “correct” analogy. We shall elaborate this 
conception of metaphysics more exactly toward the end of this paper. 

If one even agreed that the place of logic and reason in science had 
not lost in importance in our time, a great many people, including even 
scientists and philosophers, would claim that the actual advance of 
science had been promoted not by logic and reason, but rather by 
intuition and metaphysics. Logic, so this argument goes, is useful only 
for systematizing scientific knowledge and statements that are already 
known; it can never be of any help in finding new statements and, still 
less, fundamentally new theories. This assertion has been repeated 
Again and again, but it cannot stand a critical test. If we include in 


290 



logic arid mefaphysics in modern science 


logic not only syntax of language but also the theory of meaning, 
semantics, one could easily make a good case for the assertion that 
the “experimental” theory of meaning, which has been advocated by 
pragmatists and logical empiricists, is the very basis of twentieth- 
century physics. Not only does it provide the method of presenting this 
physics systematically, but one can also point out that the authors of 
this new physics made explicit use of this theory of meaning. This 
theory was one of the historical roots of the new physical theories. 

It is hardly necessary to stress the fact that this theory of meaning 
guided Einstein in his “restricted” theory of relativity as well as in his 
general theory. When Einstein in 1905 introduced his interpretation of 
the Lorentz transformation which was the essence of the special theory 
of relativity, he pointed out that the statement “two events at a spatial 
distance take place at one and the same time” cannot be used to derive 
any observable fact. Therefore, this statement cannot be a part of any 
physical theory. Einstein clearly understood that this statement needs 
in addition a semantic rule or operational definition. In the develop- 
ment of the general theory we again have statements of the type “the 
rotation of this body is responsible for a centrifugal force.” Tliis state- 
ment is regarded as a legitimate statement in Newton’s mechanics. 
But Mach pointed out that such a statement cannot be used to derive 
observable facts because it does not contain any rule by which one can 
check by observations whether a body is rotating. This remark was one 
root of the general theory and the theory of gravitation. This correla- 
tion between the theory of meaning and Einstein’s relativity was fully 
appreciated by Bridgman.^ 

It is perhaps less generally known that the development of quantum 
mechanics also has its historical root in the empirical theory of mean- 
ing. The decisive turning point in the history of quantum mechanics 
was Heisenberg’s paper* of 1925. Until this date the state of the 
quantum theory was characterized by Bohr’s theory of the hydrogen 
atom, by the Keplerean orbits of the electrons around the nucleus 
which obeyed the Newtonian laws with certain restrictions. Bohr’s 

^ P. W. Bridgman, The Logic of Modem Physics (New York: Macmillan, 19271. 

* W. Heisenberg, Zeitschrift fur Fhysik 33, 879 (1925). 

^ 291 



modern science and its philosophy 


theory superimposed upon Newton’s laws the quantum laws, accord- 
ing to which only some specific orbits could be performed without 
giving rise to a radiation which according to classical physics would 
destroy the orbit. This pre-1925 state of quantum theory can be 
described as a “Newtonian mechanics patched up by quantum laws.” 
Heisenberg, in the paper mentioned, was the first to replace Newtonian 
mechanics, in his application to the movement of electrons, by a com- 
pletely new physical theory which became known under the name 
“quantum mechanics.” His starting point was exactly the experimental 
theory of meaning. He says: 

In this paper I am going to attempt to find foundations for a mechanics 
of quantum theory. This mechanics is based exclusively on relations between 
quantities that are observable in principle. . . . As it is well known, a very 
relevant objection can be raised against the formal rules that are used in 
quantum theory for the computation of observable quantities (e.g., the 
energy of the hydrogen atom). These rules of computation contain as an 
essential part relations between quantities that are unobservable in principle 
(e.g., the position and the period of an electron). Therefore, those rules are 
without any intuitive foundation if one does not expect that those quantities 
which are at present unobservable will eventually be accessible to experi- 
mental observations. . . . Under these circumstances it seems advisable to 
make the attempt to bufld up a quantum mechanics that is analogous to the 
classical mechanics but in which only relations between observable quantities 
occur. 

Accordingly, Heisenberg introduced not the position of the electron 
but the Fourier coefiBcients of the radiation that is emitted by the atom 
as a result of the Keplerean motions of the electron. These Fourier 
coefiBcients developed later into Heisenberg’s matrices and Schrodin- 
ger’s wave function. 

Hence, the heuristic value of the “experimental theory of meaning” 
is proved by the actual history of twentieth-century physics. It turned 
out later that the actual formulation of the principles of the new 
physics could not be achieved in this direct and simple way. It has not 
been sufiBcient to use only observable quantities as terms in these prin- 
ciples. One had to proceed in a more indirect and complex way. This 
Uks been the case in relativity as well as in quantum mechanics. 


292 



logic and metaphysics in modern science 


If vre consider, as an example, Einstein’s theory of gravitation, the 
principal part of this theory is the diflFerential equations of the gravita- 
tional field. They contain mathematical symbols: the four general 
coordinates in the space-time continuum, the ten potentials of the 
general gravitational field, etc. One cannot lay down practical semantic 
rules for these symbols as they stand in the general field equations. But 
by mathematical conclusions we can derive results from these field 
equations which can be translated by means of feasible semantic rules 
into descriptions of actually observable facts. If, for example, we derive 
the bending of light rays in the gravitational field of the sun, we obtain 
statements in which the general coordinates in the space-time con- 
tinuum can be connected in a clear-cut way with the spatial and tem- 
poral distances which are measured by our traditional ways of measur- 
ing length and time intervals. But such a connection cannot be laid 
down in full generality. If, for example, we consider the statement “one 
and the same factual event can be described by different sets of values 
of the general coordinates in the four-dimensional continuum,” we can 
hardly lay down practical semantic rules by which the operational 
meaning of this statement can be defined in a simple way. 

P. W. Bridgman maintained, therefore, that Einstein’s general 
theory of relativity does not fulfill the requirements put forward by the 
experimental or operational theory of meaning. According to these re- 
quirements, one must give explicit operational definitions of all terms 
that occur in the general principles of a theory. While, according to 
Bridgman, the restricted theory of relativity was a brilliant example 
of the use of operational definitions, he thinks that the general theory 
has violated the requirement for such definitions.® There is no doubt 
that on the way from the restricted theory of relativity to the general 
theory the structure of a physical theory, as envisaged by Einstein, has 
changed in a noticeable degree.* 

The restricted theory of relativity nearly fulfilled what has been 

®P. W. Bridgman, op. cU.; The Nature of Physical Theory (Princeton Univer- 
sity Press, 1936). 

* Albert Einstein: Philosopher-Scientist (vol. 7 of The Library of Lining Phi- 
losophers, P. A. Schilpp, ed.; Evanston and Chicago: Northwestern Universit}^ 
1949). 



modern science and its philosophy 


called Mach’s “positivistic” requirement, according to which all prin- 
ciples of physics should be formulated by using only observable 
quantities as terms. The general theory made use of this requirement 
only as a heuristic principle, as a hint of how to build up the system 
of fundamental principles. These principles themselves, however, ful- 
filled the “positivistic” requirement only in an indirect way. Einstein 
replaced it, consciously and deliberately, by a weaker requirement: it 
was merely required that from these principles mathematical con- 
clusions could be drawn that were connected by semantic rules with 
statements about observable facts. 

Albert Einstein, in his Herbert Spencer Lecture given at Oxford in 
1933, speaks about this change in the way in which the abstract prin- 
ciples of physics are connected with the observable facts. This boils 
down to a change of the place where the semantic rules or operational 
definitions are attached to the abstract principles. Einstein speaks of 
“the ever-widening logical gap between the basic concepts and laws on 
the one side and the consequences to be correlated with our experiences 
on the other.” He insists merely on the requirement that some results or 
propositions in the system be connected with observational statements 
by means of semantic rules. In his “Remarks on Bertrand Russell’s 
Theory of Knowledge,” ® Einstein says: 


In order that thinking might not degenerate into “metaphysics” or into 
empty talk, it is only necessary that enough propositions of the conceptual 
system be firmly enough connected with sensory experiences. 




According to this new conception, the sentences that have to be 
connected with sense observations by semantic rules are no longer the 
general abstract principles (such as the law of conservation of energy) 
but some special conclusions drawn from these principles. The “posi- 
tivistic requirement” now means that there must be some consequences 
of the general principles which can be translated into statements about 
sense observations. The general principles themselves are the product 


“ The Philosophy of Bertrand Bussell (vol. 5 of The Library of Living Philoso- 
phers, P. A. Schilpp, ed.; Evanston and Chicago: Northwestern University, 1944), 
p. 289. 


294 



logic and metaphysics In modern science 


of mathematical and logical imagination which has to be checked by 
applying the “positivistic” or “operational” requirement. 

The nature of a scientific theory in this sense can be understood 
even more precisely if we consider the new “unified field theories” pro- 
posed by Einstein and Schrodinger. A “unified field of force” is to be 
constructed which contains the gravitational, the electromagnetic and 
the nuclear field as special cases. Schrodinger, for example, introduces 
sixty-four symbols which are the components of this unified field. He 
does not lay down any semantic rules for these sixty-four quantities. 
But if this theory is to be of any scientific value, he has to assume, as a 
matter of course, that some special relations can be mathematically de- 
rived from the principles which can be connected with observable 
facts. He hoped, for example, that it could be derived that a rotating 
mass which obviously produces a rotating gravitational field would 
entail a magnetic field. One would be able to account in this way for 
terrestrial magnetism. 

From a psychological viewpoint Einstein describes this way of 
producing theories by free imagination in a letter to the French 
mathematician Hadamard.® According to this new conception, it is 
true that physical theories are the product of free imagination, if we 
take the word “free” with a grain of salt. But it must not be concluded 
that these theories are products of metaphysics. For these theories are 
subjected to the operational or experimental criterion of meaning, 
though in a more indirect and complex way. The criterion of truth 
remains ultimately with the checking by sense observations, as the 
older “positivists” claimed. But we know now that this checking is a 
more complex process than it was believed by men like Comte and 
Mach to be. 

If we applied the name "metaphysics” to a system of statements the 
“truth” of which is judged according to the experimental criterion of ^ 
meaning, there would be no distinction between science and meta- 
physics. We have, therefore, to reserve the word “metaphysics” as a 
characteristic of systems the truth of which is decided on other grounds. 

* J. Hadamard, An Essay on the Psychology of Invention in the Mathematical 
Field (Princeton University Press, 1045). • 

* 295 



modern science and its philosophy 


In metaphysics a statement or a system of statements is regarded as 
“true” if our common sense imderstands the validity of the principles 
immediately without having to draw long chains of conclusions from 
these principles and without checking some of these conclusions against 
our observations. 

Such a metaphysical interpretation of twentieth-century physics 
was given, for instance, by Eddington.’ He claimed that the validity of 
our physical theories can be demonstrated by what he called “epistemo- 
logical arguments.” This meant in his language that the principles of 
physics have to meet some requirements emerging from common sense. 
And these requirements are sufficient to determine our physics to such 
a degree that even the number of the electrons in the whole universe 
can be derived mathematically. Eddington’s argument is actually 
“metaphysical.” We could call it “epistemological” only if we regarded 
epistemology as a part of metaphysics. The point is that Eddington 
derives his system of physics from everyday experience and not from 
scientific experiments which are needed to check the results of a long 
chain of conclusions from the principles. The requirements of “com- 
mon sense” actually are that these principles should be a convenient 
description of our everyday experience. 

Certainly, men like Einstein and Schrodinger advanced their 
principles by following some requirements of simplicity or beauty 
which may also be regarded as requirements of common sense. But 
they would never claim that the validity of the principles could be 
proved without checking the conclusions drawn from these principles 
by physical experiments. 

If we want to compare the respective places of logic and meta- 
physics in the actual advancement of science, we can point out two 
ways in which logic has been instrumental in the advance of twentieth- 
. century science. We have already described the heuristic value of the 
experimental theory of meaning which played a decisive role in the 
rise of relativity as well as of quantum theory. But logic has also played 
in another way a guiding role in the advance of twentieth-century 

’ A. S. Eddington, The Philosophy of Physical Science (New York: Macmillan, 
1639 ). 


296 



logic and metaphysics in modern science 

science. And this way is connected with formal logic, or what we may 
call ‘logical imagination.” Einstein, in particular, has described re- 
peatedly how the building of formal systems of symbols, the successive 
demolition, rebuilding and alteration of these symbolic structures, bas 
had a great bearing upon the advance of new physical theories. In 
his letter to Hadamard, Einstein gives a psychological description of 
his creative work. He insists that the essential part in creative thinking ✓ 
is the free play with symbols. In his Herbert Spencer Lecture (1933), 
Einstein says: 

Experience, of course, remains the sole criterion for the serviceability 
of mathematical construction for physics, but the truly creative principle 
resides in mathematics. 

This means, obviously, that the creative process in theoretical 
physics consists, in some important ways, in the creation of symbolic 
or formal systems by a kind of “logical imagination.” Among the sys- 
tems created in this way, experience is responsible for the natural 
selection that determines which system is the fittest for survival and 
which has to be dropped. 

In order to appreciate the place of logic, we have to consider the 
shift in the conception of a physical theory that has developed from the 
times of Mach to the times of Einstein. The operational definitions or 
semantic rules are now no longer applied to the general principles 
themselves but to some conclusions drawn from them. However, this 
distinction should not be overstressed. It would be erroneous to be- 
lieve that men like Ernst Mach actually believed that all principles of 
physics were direct descriptions of experimental facts. In his paper on 
the role of comparison in physics, Mach distinguished very precisely 
between “direct description” and “indirect description.” The latter, he 
said, is also called “physical theory.” He gives as an example the wave 
theory of light. In this theory there is certainly no explicit operational 
definition of the “light vector.” 

As for the heuristic value of metaphysics, we may quote one of the 
most prominent contemporary advocates of metaphysics, Jacques 
Maritain. He says bluntly: 


297 



modern science and its philosophy 


It is true that metaphysics brings no harvest in the field of experimental 
science ... Its heuristic value, as the phrase goes, is nil . . . 

It cannot be of any help in promoting scientific research. He continues: 

This universe in which metaphysics issues ... is not intelligible by 
dianoetic or experimental means, it is not connatural to our powers of knowl- 
edge, it is only intelligible to us by analogy. 

This means in the usual language of the scientist: In metaphysics we 
do not use either logical (dianoetic) or experimental argument, but 
we interpret the principles of science in a metaphoric language. Meta- 
physics attempts to interpret the general principles of science in a way 
that is plausible to our common sense. The principles then become 
analogous to the laws of everyday experience; they duplicate them, but 
on a higher level. The law of conservation of energy becomes meta- 
physically plausible because it is bolstered up by the well-known ex- 
perience that objects of the physical world (stones, animals, etc.) 
cannot disappear. We transfer the conception of “disappearing” to an 
“object” called “energy” which is certainly not a physical object in 
the sense that a table is such an object. Therefore “disappearance of 
energy” can only be understood by “analogy” with the disappearance 
of a stone. According to Thomistic metaphysics, in an expression like 
"the being of God” or “the being of a spirit” the word ‘Treing” does not 
mean the same thing as it does in “the being of a stone.” The meaning 
of “being” on these “higher levels” is only understandable by analogy, 
not by any direct operational definition or semantic rule. 

I would not even go as far as Maritain in denying any heuristic 
value in metaphysics. The description by analogies can occasionally 
be of some psychologic value in setting up new principles. This was 
even recognized by so staunch an opponent of metaphysics as Ernst 
Mach in his paper on the role of comparison in physics. 

But if we analyze a little further the nature of the actual metaphysi- 
cal interpretations of physics, we soon notice one characteristic of them 
that can easily become prohibitive to any future advance of physics. 

I shall not now go into an elaborate discussion of this nature of 
ifletaphysics. I shall restrict myself to showing by some examples that 


298 



logic and metaphysics in modern science 


philosopheis who went in for both metaphysics and science frequently 
pointed out this characteristic of metaphysics. They stressed the point 
that metaphysics is an attempt to interpret the general principles of 
physics in terms of the language of our everyday life. In this way it is 
possible to muster the support of our everyday experience to make 
those principles plausible. 

A very characteristic example is the French philosopher E. le Roy. 
His goal was to prove on the basis of contemporary science that room is 
left for metaphysics and even for a metaphysical conception of religion. 
He especially liked to make use of the ideas that the famous French 
scientist, Henri Poincar4, had advanced about the logical status of 
science. Just at the turn of the century (1899) Le Roy published in 
the Revue de M^aphysique et de Morale a paper in which he says: 

Science departs from common sense and does not join it in its develop- 
ment as science proper. Thus science by itself does not close the cycle of 
knowledge and does not realize the unity of knowledge. Science needs there- 
fore a prolongation and this will be philosophy ... in one way science 
itself is a prolongation of common sense. 

Common sense, science, philosophy, common sense form a cycle. 

The term “philosophy” refers here, of course, to the same thing as 
“metaphysics.” In these words we see clearly the place assigned to 
metaphysics. Science departs from common sense by using a different 
conceptual scheme; words are used in a different way. Science even 
introduces expressions that have no meaning in the language of com- 
mon sense. This becomes particularly clear if we consider the language 
of twentieth-century physics. The theory of relativity uses expressions 
like "one and the same object has different lengths relative to different 
systems of reference,” and quantum mechanics uses expressions like 
“if the position of a particle is a definite one, its velocity is always 
indeterminate.” 

According to Le Roy’s cycle, there are two ways of connecting 
science with common sense: the direct, which we may call the “scien- 
tific” one, and a second, which connects science with common sense 
by means of metaphysics. In the scientific way, the statements of 
science are interpreted by means of operational definitions as state- 


299 



modern science and its philosophy 


ments about observable facts, and every observable fact in physics can 
be expressed in the language of common-sense experience. 

Einstein’s statement that a rigid body has different lengths with 
respect to different systems of reference can be connected with com- 
mon-sense statements in two ways. We can describe directly the way in 
which the length of one and the same rigid body can be measured by 
putting end to end yardsticks that have different velocities relative to 
the body. In contrast to this scientific connection, the metaphysical 
interpretation would be: the statement that “one and the same length 
is estimated differently by different observers” reminds us of the ex- 
perience of everyday life that one and the same length is estimated 
differently by different observers. This “subjectivity” of every judgment 
about length seems analogous to Einstein’s contention that the length 
depends on the .system of reference. Therefore, Einstein’s statement is 
interpreted as claiming the “subjectivity” of human statements about 
length. This is again in line with some statements of idealistic 
philosophy. 

We also find similar views in writings of other philosophers of 
scientific background. I mention A. N. Whitehead, C. S. Peirce, and 
in particular, a Thomistic philosopher, H. V, Gill. He regards the meta- 
physical interpretation, as many people do, as a result of “intuition.” 
But he realizes somehow that it is actually an application of common- 
sense judgment to the principles of science. He noticed that the 
scientists (he calls them “specialists”) will not do as well in finding 
these interpretations as the “man in the street” who relies on common 
sense only. Gill says; ® 

The fact that a few “specialists” call in question some intuition generally 
accepted by men does not furnish a valid reason for doubting its truth. The 
specialist is indeed perhaps the one whose view on first principles should be 
taken most cautiously. 

The metaphysical interpretation is actually a particular kind of 
^ semantic approach; it is a translation into common-sense language. 
We follow Gharles Morris’s excellent analysis of the "metaphysical 

*■ * H. V. Gill, Fact and Fiction in Modem Science (Dublin: Gill, 1943), p. 21. 



logic and metaphysics in modern science 


discourse.” ’ It is a “formative discourse” like the mathematical, logical, 
grammatical and rhetorical discourse. A criterion of truth (similar to 
the criterion of “scientific truth”) cannot be applied. The metaphysical 
discourse plays a role in organizing human behavior and has, therefore, ^ 
“significance.” The present paper means to be more specific and to 
describe the language used in metaphysics as the result of an attempt 
to interpret the general laws of science by using common-sense ex- 
pressions. The “materialistic” and “idealistic” interpretations of science 
owe their appeal to the common-sense meaning of the words “matter” ^ 
and “mind” and not to their scientific meaning, which can hardly be 
stated precisely. 

From these considerations we can easily derive a sound judgment 
about the role of metaphysics in the actual advance of science. What 
we call in a vague way “common sense” is actually an older system of 
science which was dropped because new discoveries demanded a new 
conceptual scheme, a new language of science. Therefore the attempt 
to interpret scientific principles by “common sense” means actually an 
attempt to formulate our actual science by the conceptual scheme that 
was adequate to an older stage of science, now abandoned. 

According to these considerations, one can easily estimate what the 
role of metaphysics in the advance of science has been. To believe that 
some “metaphysical interpretation” may tell us the “truth” about the 
“real world” means in practice to believe that the conceptual scheme of 
some older stage of science is necessarily the scheme to be used for all 
the future. This belief is, certainly, in a way stimulating: it encourages 
the scientists in their attempt to stick to a unified scheme into which 
every new discovery has to be fitted or perhaps even squeezed. To 
achieve a unified scheme in all fields of physics is certainly a goal 
that has been occasionally of great heuristic value. The most prominent 
example is, I think, the attempt to interpret all physical phenomena by 
Newton’s laws of motion. This attempt has led to great successes in 
optics; we owe to it the corpuscular theory of light as well as the wave 
theory. We owe to it almost all atomistic theories in their first stages. 

*C. Morris, Signs, Language and Behavior (New York: Prentice-Hall, 1946), 
pp. 175 ff. 


301 



modern science and its philosophy 


But toward the end of the nineteenth century the physicists came more 
and more to recognize that there are phenomena which can be fitted 
into this “mechanistic” pattern only very artificially and incompletely. 
Hence the heuristic value of this “mechanistic” goal faded as time 
went on. Nevertheless the belief remained that only physical theories 
which can be derived from Newton s laws of motion satisfy human 
desire for “understanding” nature. In this stage this belief became a 
purely metaphysical creed. Some maintained, for example, that Ein- 
stein s modification of Newton’s laws had to be rejected for meta- 
physical reasons. This actually means only that Einstein’s mechanics 
cannot be derived from Newton’s laws of motion. It is a historic fact 
that a great many physicists preferred to say that they rejected rela- 
tivistic mechanics not for metaphysical reasons but for reasons of 
“common sense.” Both types of reason for rejecting new physical 
theories, as the argument in this paper shows, are really one and the 
same thing expressed in two different ways. 

The rationalistic metaphysician rejected new theories on die 
ground of “reason,” the empiricist-metaphysician rejected them on the 
basis of “common sense.” In my view, both types of rejection have a 
common origin: the belief in the interpretation of new theories by using 
the language of older theories. 

Examples are abundant. I mention only cases in which “common 
sense” prevented the acceptance of new physical theories, because the 
scientists are more easily caught by “common sense” than by avowed 
metaphysics. 

The father of empiristic philosophy, Francis Bacon, rejected the 
Copemican theory for not being in agreement with common sense; 
the leader of nineteenth-century British empiricism, Herbert Spencer, 
argued that the total mass of a system of material bodies cannot de- 
pend on their distribution in space. August Comte, the father of 
“positive philosophy,” predicted that no mathematical theory of 
chemical phenomena will ever be advanced because our common sense 
tells us that the chemical processes are fundamentally different from 
physical processes. If we consider to what degree all these predictions 
have been refuted by the actual advance of science, we can learn two 


302 



logic and metaphysics in modern science 


things: metaphysics has very often been an obstacle to the advance of 
science, and second, if we hear today that biology will never become 
a science in the sense that mathematical physics is, or that sociology 
can never use scientific methods, we shall hesitate in maintaining a 
smug belief in these assertions. 


303 




BIBLIOGRAPHICAL NOTE 


The essays in this volume have appeared in the following journals: 

1. “Kausalgesetz und Erfahrung,” Ostwald’s Annden der NaUirphtbso- 
phie 6, 443 (Leipzig, 1907). 

2. “Die Bedeutung der physikalischen Erkenntnistheorie Machs fur das 
Geistesleben der Gegenwart,” Naturtcissemchaften 5, 65 (Berlin, 1917). 

3. “Ernst Mach— the centenary of his birth,” Erkerintnis 7, 247 (The 
Hague, 1938). 

4. “Was bedeulen die gegenwartigen physikalischen Theorien fiir die 
allgemeine Erkermtnislehre?” Efkenntnis 1, 126 (Leipzig, 1930). 

5. “La physique contemporaine manifeste-t-elle une tendence a r6int6grer 
un 616ment psychique?” Revue de synthdse 8, 133 (Paris, 1934). 

6 and 10. "The mechanical versus the mathematical conception of nature,” 
Philosophy of Science 4, 41 (1937). 

7. “Modern physics and common sense,” Scripta MathemaHca 6, No. 1 
(1939). 

8. “Die philosophischen Missdeutungen der Quantentheorie,” Erkennt- 
nis (Leipzig, 1936). 

9. “Bemerkungen zu E. Cassirer: Determinismus und Indeterminismus 
in der modemen Physik,” Theoria 4, 70 (Gbteborg, 1938). 

11. “Logisierender Empirismus in der Philosophie der U.S.S.R.,” Actes du 
Congris International de PhUosophie Sdentifique (Paris, 1936). 

12. “Why do scientists and philosophers so often disagree about the 
merits of a new theory?” Reviews of Modem Physics 13, 171 ( 1941) . 

13. “The philosophical meaning of the Copernican revolution,” Proceed- 
ings of the American PhUosophicd Society 87, 381 (1944). 

14. "The place of the philosophy of science in the curriculum of the 
physics student,” American Journal of Physics 15, 202 (1947). 

15. “Science teaching and the humanities,” Etc.: A Review of General 
Semantics 4, 3 ( 1946) . 

16. “The place of logic and metaphysics in the advancement of modem 
science,” Philosophy of Science 15, 275 (1947). 

Chapters 1, 2, 3, 4, 5, 8, 9, and 11 also appeared in Between Physics 
and Philosophy (Harvard University Press, 1941). 




INDEX 


Acceleiation, operational meaning of, 
236, 245 

Accuracy, arbitrary, 117 
Action, procedures of, 5; existence of 
quantum of, 108, 110; spontaneity of, 
162; objective observation of human, 
166; at a distance, 211 
Adler, Mortimer J., 271 
Age of Enlightenment, 72, 74, 280 
Age of Galiko, 122 
Age of Newton, 122 
Agnosticism, 14; Mach accused of, 17 
Agreement, of physical theory with ob- 
servations and principles, 210 
Alembert. Jean Le Rond d’, 75 
America, 216 

Analogy, representation by, 150; to or- 
ganism or mechanism, 252; tradi- 
tional, 253; universe intelligible by, 
298 

Analysis, logico-empirical, 245, 253; 
semantic, 245, 249; socio-psychologic, 
249; pragmatic, 249; historical, 278 
Analysis of Sensations, The (Mach), 
69, 81 

Angels, behavior of, 271 
Animism, rejection of medieval, 140; in 
Newton’s conception of force, 224 
Animistic science, return to, 194 
Anschaulich, 147 
Anschamng, 147, 197 
Approach, coherent, 250; socio-psycho- 
logic, 250; semantic, 300 
Approximations, successive, 118 
Aquinas, St. Thomas, 219, 220, 221, 
223, 251, 271, 284, 286 
Arbitrariness, 266 

Argument, philosophic and historical, 
276; epistemological, 296 
Aristotelianism, 73, 281; fundamentalist 
and modernist attitudes, 23-24 


Aristotle, 23, 37, 103, 115, 218, 284, 
287 

Arrangements, complementary, 163 
Art, as personal experience, 42 
Assertion, idealistic, 112 
Assignment, rules of, 120 
Astronomical instruments, 146 
Astronomy, 271; Greek, 23; no true, 225 
Atom(s), 67; concept of, a fiction, 43; 
electrons and, 66, 116; no longer rigid 
lumps of reality, 188; dematerializa- 
tion, 190 

Atomic bomb, 262 
Atomic energy, 254 
Atomic processes, laws of, 125 
Atomism, 143; mechanical, 58; aversion 
of Mach and Duhem for, 140 
Atomistics, 65, 76, 116, 117; Mach's 
opposition to, 70 
Attitude, complacent, 279 
Austria, 18, 26 
Avenoes, 221, 224 

Axioms, 242; arbitrary, 12; of coordina- 
tion, 31; of geometry, 134, 244; of 
parallels, 134; self-evident, 254; of 
Newtonian mechanics, 270. See dso 
Structural system 
Ayer, A. J., 48 


Background, cultural, philosophy of sci- 
ence and, 42 

Bacon, Francis, 209, 210, 232, 302 
Balmer series, 107 
Bavink, Bernard, 123, 125, 187-189 
Beast of prey, philosophical, 76 
Beauty, requirements of, 296 
Becker, Carl, 286-289 
Behaviorism, 166 

Being, kinds of, 37 , ^ 

Bergson, Henri, 95, 101, 115 


307 


1 



index 


Berlceley, Bishop George, 85, 193, 199, 

201, 210 

Bible, cntenon of fruits, 64, as scaentific 
textbook, 74 
Biological advantage, 65 
Biological-pragmatic approach, to sci- 
entific conception of nature, 110 
Biology, 84, concepts of, 37, evolution- 
ary, 103, antimechanical, 125, vital- 
istic, 159, understanding of, in terms 
of physics, 165, Bohr’s thought m, 
170, application of complementary 
prmciple to, 180 
Biophysics, 273 
Blumberg, A , 38 

Bodies, properties of, 56, 58, movement 
of tvo, 104, rigid, 107, 136, of aver- 
age size with moderate velocities, 
131, of everyday experience, 153, ter- 
restrial and celesbal, 221, moving, 
contraction of, 226, problem of three, 
235, material, existence of, 270 
Body, "real” length of, 96, test, 117, 
gulf between mmd and, 122, physical 
causalitv m human, 166 
Bohr, Niels, 179, principle of comple- 
mentarit), 15^159, 163, 165, 166, 
170, 182, 240, 246 255, 275, theory 
of hydrogen atom, 291 
Boll, Marcel, 48 

Boltzmann, Ludwig, 141, 142, 150, 250 
Bolyai. J4nos, 270 
Bourgeoisie, 200 
Bndges, logical, 173 
Bridgman, P W , 20, 246, 250, 254, 
256, 293, theory of meaning, 44-45, 
224, 225, 244, 247, 253 
Brvson, Lyman, 52 
Burtt, 224 


Calculus, differential, 135, 194 
Cambridge, England, 49 
Cambridge, Massachusetts, 49 
Canossa, scientist’s trip to, 92 
Carnap, Rudolf, 37, 38, 41, 44, 45, 48, 
84, 85, 161, 166, 171, 202, 256, 
philosophical system, 33-36, 86, 111- 
112, 121, 152, 162 
^artesian system, 236-238 
Cartesians, 132 


Cash value, meaning as, 32, 33 
Cassuer, Ernst, 24, 172, 17^183 pasnm 
Categories, theoiy of, 9 
Catliolic Church See Homan Church 
Catholic colleges, 256 
Catholicism, Duhem’s belief m, 16 
Causality, 112, concept of, 115, 255, 
discussions of, 207, framework of, 
120, law of, 10-11, 53-60, 175, 176, 
270, as arbitrary convenbon, 14, es- 
tablishment of a terminology, 57, 
Kant’s formulabon, 177, validity or 
nonvalidity, 117—118, physical, in hu- 
man body, 166, strict, 119 
Causes, mechanical and orgamc, 4, real, 
46, my sbcal vital, 96 
Celesbal bodies, 221 
Cenbal Europe, movement toward sci- 
entific world conception, 26, 44 
Centrifugal force, 225 
Changes, speculative, 113 
Chemist, professional, 232 
Chemistry, classes m, 232 
Chicago, University of, 48 
Chnstoffel, Elwin Bruno, 134, 135, 136 
Churchman, C West, 42 
Circles, eccenbic, 221 
Civilizabon, technical, 122, cause of, 
260 

Clarity, lack of, in termmology, 164 
Clirke, Samuel, 132 
Classes, science, 243 
Clavius, C , 208 
Clergymen, 263 
Climate of opinion, 287 
Clockwork, understandmg of, 211 
Coefficients, Fourier, 292 
Cognition, 113, Schlick’s theory of, 29- 
30, 42 

Cohen, Hermann, 194 
Colonng, emofaonal, 173 
Colors, Goethe’s doebme of, 70 
Columbia University, 48 
Columbus, Christopher, 101, 108, 216 
Common sense, 209, 232, 240, 252, 279, 
289, m mterpretation of experiments 
97, contradicbon to, 144, accordance 
with, 156, and pnnciples of science, 
289, 300, criteria of, 290, require- 
ments of, 296, departure of science 
from, 299, m mteipretabon of laws 


308 



index 


of science, 301; an older system of 
science, 301 

Common-sense conception, untutored 
man’s, 184 

Common-sense method, Galileo’s, 286 
Common-sense statements, 300 
Communism, 191, 263. See also Soviet 
Union 

Communist revolution, 201 
Comparison, role of, in physics, 298 
Complementarity concept, 162-166 
passim, 182, 240, 246, 255, 275; lack 
of argument for free will, 170 
Complexes of perceptions, 80 
Compton effect, 117, 169 
Computations, matliematical, 267 
Comte, Auguste, 14, 290; positive phi- 
losophy of, 9, 12, 166, 256, 302; defi- 
nition of philosophy, 37 
Conant, James B., 50 
Concepts, geometric, 12-13; auxiliary, 
18, 67, 73, 74. 99, 174-175; observa- 
tional and auxiliary, 27-28; abstract 
and of observational, 30-31; meta- 
physical, 37, 83; theologic, 74; eight- 
eenth-century, 76; constitution of, 
111; mathematical, 135 
Conclusion, logical, 242, 289 
Confirmation, degree of, 28, 288; ex- 
perimental, mathematical proof and, 
234 

Conscience, voice of, 277 
Consequences, agreement of, with ob- 
servations, 210 

“Conservation of Energy’’ (Helmholtz), 
213 

Conventionalism, 11, 100-101, 176 
Conventionalistic assertion, 176 
Conventions, laws of science as, 10- 
11 

Cooperation, organization of interna- 
tional, 49 
Coordinates, 236 

Coordination, axioms of, 31; relations of, 
44; rules of, 245 
Copenhagen, 49, 179 
Copemican world system, 90-91, 216- 
227; conflict with Roman Church, 16, 
47, 218, 226, 232, 268; attitude of 
scientists and philosophers toward, 
207 


Copernicus, Nicholas, 63, 73, 90, 144, 
210, 214, 216-227 passim, 268, 284 
Corpuscular theory, 246 
Correlations, epistemological, 244 
Correspondence, unique, 29, 105; 

Schlick’s theory of cognition as es- 
tablishment of, 29-30 
Cosmological problems, 237 
Courses, 232; elementary science, 249; 

survey, for advanced students, M4 
Creeds, metaphysical, 46; political and 
religious, 46, 258, 281 
Criminal, responsibility of, 167 
Criticism, immanent, 172; scholastic, 
173 

Critique of Pure Reasort (Kant), 57 
Crystal lattice, 113 
Curriculum, college, 51 
Czechoslovakia, 18, 26, 39, 45 

Defeatism, attitude of, 269 
Definitions, operation^, 20, 22, 28, 44, 
244, 253, 278, 291; conventional, 53, 
54; disguised, 57, 100 
Demagogues, 266 
Demiurge, creations of, 133 
Democracy, fight against liberalism and, 
280 

Democritus, 58, 196, 211 
Descartes, Ren4, 211; issues between 
Newton and, 279 

Description(s), Mach’s view of explana- 
tion and, 6-7; Duhem’s theory, 16; 
simplest, 65; theories as pure, 143; 
space-time and causal, 164; comple- 
mentary, 164, 167; classical and 
quantum-mechanical, 170; of physical 
devices, 267; direct and indirect, 297 
Determinism, 46; strict, 96; philosophic, 
176; concept of, 255; overthrow of 
physical, 282 
Deterministic laws, 176 
Deterministic postulate, 177 
Deviations, idealistic, 203 
Dewey, John, 256 
Dialectical Materialism, 191 
Dialectics, trivialization of, 202 
Diamat. See Materialism, dialectical 
Differential equations, 130, 143 
Dingier, Hugo, 101 
Dirac, P. A. M., 179 


309 



Index 


Discourse, metaphysical, 46, 290, 300- 
301 

Disputes, of metaphysical character, 152 
Dnieprostroy, 201 

Doctrines, traditional, and school philos- 
ophy, 40; medieval, 122; physical, 
philosophic and religious implications, 
263 

Driesch, Hans, 53-54, 57-60 passim 
Du Bois-Reymond, Emil, 92-93, 97, 213 
Duhem, Pierre, 14, 17, 20, 30, 49, 100, 
140, 178, 257; physical theories, 15- 
17, 21, 25, 04 


Earth, mobility of, 222; influence of ro- 
tation of, 237 
Eccentrics, 219, 222 
Economy, Mach principle of, 63; capi- 
talist, 126 

Eddington, Sir Arthur Stanley, 130, 243, 
257, 296 

Education, general, 233, 261, 262, 271,- 
specialization versus general, 261, 273 
Ehrenhaft, F., 71 

Eighteenth century, 212, 281; Mach’s 
nostalgia for, 72; spirit of, 224 
Einstein, Albert, 124, 136, 180, 210, 
234, 263, 295, 296; on law of causal- 
ity, 10-11; relativity theory, 18-21, 
26, 68, 73-74, 127, 131, 225, 229, 
275, 291; lecture on “Geometry and 
Experience,” 22-23; integration into 
new positivism, 28, 44, 48; physics 
of, 130; theory of gravitation, 134- 
135, 227; conflict with Newtonians, 
268, 270; Herbert Spencer Lecture, 
294, 297 

Einstein time scale, 113 
Einsteinian revolution, 216, 225 
Electric field, measuring intensity of, 
117 

Electric forces, as elements of reality, 27 
Electric intensities, 59, 115 
Electrical phenomena, true cognition of, 
106 

Electricity, 58; nature of, 93, 98 
Electromagnetic field, Maxwell's, 140, 
239 

Electromagnetic mass, 161 
Electrons, 67; atoms and, 60; coinci- 


dence between, 116; light diffracted 
or scattered by, 116, 118; personifica- 
tion of, 118; positions and velocities 
of, 118 

Elements, intuitional, 27 
Emerson, Ralph Waldo, 228, 232 
Empathic understanding, 197 
Empirical Philosophy, Society for, 40 
Empirical tests, 147 

Empiricism, 251; logical, 16, 45-49, 
85, 86, 173, 175, 180, 197, 203; Eng- 
lish, 195; positivistic, superiority of 
mechanism, 196; crawling, 198; ad- 
vocate of, 209; crude, 265 
Empiricist, pure, 97 
Empiricist-metaphysician, 302 
Energetics, 58, 137, 158 
Energy, conservation of, 53, 164, 220, 
266; matter and, 66; minimum of dif- 
ferent kinds, 71; degradation of, 123; 
concept of, 255; disappearance of, 
298 

Energy transformations, phenomena as, 
158 

Engels, Friedrich, 202, 257, 266 
England: See Great Britain 
Ennghtenment, Mach as philosopher of, 
17-18; philosophy of, 72-78; spear- 
head of, 267 
Entelechy, 59, 60 
Entities, metaphysical, 18 
Entity, Neurams avoidance of word, 
35 

Epicureans, 281; materialism of, 132 
Epicycles, 219, 221, 222 
Epistemological foundation, 70 
Episteraologist, 91 

“Epistemology of the Exact Sciences,” 
40 

Equations, field, 113, 140; differentia], 
130, 143; Newton’s, 242 
Equivalence, principles of, in Einstein’s 
theory, 18 

Erkennfnis, mouthpiece of Vienna 
group, 41 

Ernst Mach and Marxism (Valentinov), 
190 

Ernst Mach Association, 40 
Essence, Neurath’s avoidance of word, 
35 

Essential, misinterpretation of word, 35 


310 



index 


ETC.. A Review of Genet d Semantia, 
247 

Ether, matter-like, 153 Ntlocity with 
respect to the, 154, 155 
Ethics, 24, 277, eudacmonistic, and pos- 
itivistic sociology, 39, relation to 
quantum mechanics, 184, absolute 
\ allies of, 230 factois of, 255, syn- 
thesis with science, politics, and reli- 
gion, 271 

Euclid, 12-13, axioms of c oordinaboii, 
31 

rxistcncc, real, lOS, of external objects, 
268 

Lxpcncncc(s), Kant’s theory of, 7 8, 
objectnc, and mind, 9, leisoning and, 
in geometrj, 21-23, dcsiie for pei- 
sonal, 42, science as economical rcp- 
lesentation of, 84, unequisocal as- 
signment of s)mbols to, 106, concrete, 
similarity between. 111, real, 118, 
correspondence betw een symbols and, 
118-119, point, 119, everydav, 145, 
149, 152, 288, naive sense, 217, 218 
Experiments, world of physical, 243 
Experimeutum cruets, 15 
Explanation, scientific and mechanical, 
6, Duhem’s theory of, 16, mecha- 
nistic, 138-140, causal, 211, plqsical, 
226 

Explanatory \ alue, of mechanistic theo- 
ries, 139 

Exposition, emotional undertone of, 174 

ract(s), obserxed, in Mach’s theory of 
science, 11, obserx ational, Einstein’s 
tlicory in terms of, 19, 20, economical 
dtsciiption of, 20-21 xxorld of, 42 
objectixe, 80, 127, 128 statement 
about obseiyed, 242, confirmable, 
266, marriage of reason and, 286 
Factors, extrascientific, 283 
Faith, matters of, 130 
Farada) , Michael, lines of force, 63 
ratadat/s Lines of Fotce (Maxwell), 
141 

Fascism, 266, 282 
Feeling, optimistic, 42 
Feigl, H , 38 

Feudalism, return to ideas of, 282 
Fictions, mathematical, 91 


Field equations, integration of, 113, 
Maxwell’s, 139-140 
Field intensities, 115 
Finkelstein, Louis, 52 
Fislce, John, 257 
Fluid vortices, 132 
Fog, enigmatic, 229 
Force, Faraday’s lines of, 63, concept 
of, 80, 247, nature of, 93, 99 
Formal idiom, 152 

Formulas, matliematical, 46, 138, physi- 
cal meaning of, 267 

Fonnulations, sensible and meaningless, 
of problems, 95, metaphysical, 164 
Fossihzation, 208 
Foul dimensional space, 152, 158 
Fouiier coefficients of radiation, 292 
Frame of reference, 154 
Framcyxork, man’s, in cribcal idealism, 
58, rigid, of space, tune, and causal- 
ity, 120 

France, materialism in, 125 
Frankenstein’s monster, 283 
Free will, 159, 162, 166, 183, science 
and, 125, not an eipression from 
psychology, 167, belief m, 230 
Freedom, from intoxication and hypno- 
sis, 167, from external coercion, 167, 
itlucal, 183 real and pseudo, 238 
Fiege, 103 

French, peciiliai ties of, 148 
Flench Rexolution, 124, 216, 281 
Fruits, criterion of, 64, of Mach’s teach- 
ings, 68 

Galactic s\ stem, 226 
Galaxies 237 

Galileo 90, 159 186, 214, 223, 286, 
287 dialogues of, 73, trial of, 91, 
physics of, 115, 123, 138 
GaJvanometer, 107 
Gauss, Karl Friedrich, 270 
General Education Program, Harvard 
Umxersity, 50 

Generalizations, popular, 274 
Geoccntiic system, 146, 208, 218, 222, 
225 

Geodesics, 130 
Geometnc imagery, 151 
Geometric theorem, 233 • , 

Geometry, 242, 243, Euclidean, 8, 9, 


311 



index 


12-13, 21-24 paisim, 91, 114, 134, 
147, 270, non-Eudidean, 8, 9, 21- 

22, 91, 147, 233, 234, 254, 270, foun- 
dahons of, 13, 234, 239, 240, 241, 
245, reasoning and experience m 21- 

23, Rcichenbach’s conception of, 31, 
Kicnunnian, 136, iinderstanding of, 
235, axioms of, 244, logico-empirical 
analysis of, 253, niathcmatiLal, 253, 
semantic analysis of, 254 

“Geometry and CxpericiiLc” (Einstein), 
22, 245 

German Plixsital Soc.ict\. 39-10 
German), 124, Weimar expenment, 26 
unixersities of, 31, niittnalism in, 
125 Na7i. 195 
Gill, H V , 300 

God, 222, science and, 125, \xill of, 277 
Goethe, Johann Wolfgang \on, 62-63, 
doctiinc ol colors, 70, extreme phe- 
nomalist, 71 
Good life, 272 
Good xmII, 279 

Government, relations between science, 
religion and, 232 

Gravitation, 123 theory of, 135, law of, 
210, 211, 212, 214, 217, Einstein’s 
till orv of, 227 

Gravitational field, Einstein’s laws of, 18 
Great Britain, 48, 50, analytic and spec- 
ulative philosophy in, 36 
Groups, totalitarian. 34 

Iladamard, J , 295 
Hahn, Hans, 1, 31 32, 33, 34, 38 
Hardness, meisiircment of, 114 
Harvard Observatorv, 226 
Harvard Umversitv, 48, 49, 50, 250 
Havakavva, S 1 , 247 
Heavens, real motions of 222 
Hegrl, G W r , 88, 202, 284, objective 
idealism of, 201 

I Ii gtlianism, idealistic eggshell of, 203 
Heisenberg, Werner, 116, 127, 179, 291, 
292, principle of uncertainty, 168, 
240, 275 

Heliocentric theory, 144-146, 218, 225 
Helmholtz, H L F von, 134, 136, 138, 
213 

Herbert Spencer Lecture, 294, 297 
Hermann, Crete, 177, 184 


Hertz, lleiniieh, 93, 140, 141 
Heuristic principle, 294 
High frequencies, 117 
Hilbert, David, new foundations of ge- 
ometry, 13 103 
Ihllebrand, K , 195-197 
History, 276, of science, 149, courses m, 
232, of British kings, 233, laws of, 
271 

Holism, 195 
Human interest, 232 
Human society, crisis in, 136 
Huniaiiitics, 232 stienct and, 258, 261, 
281, 284, basis of education, 260, m- 
Icicst 111 , 278, instruction in, 283 
Hume, David, 12, II 242, 256, law of 
causality, 10 

Ilinghens, Christian 138 
Ilvdrogcn atom, Bohi’s tlieoiy ol, 291 
Ilvpotheses, as ‘ciffolds, 62, niinimum 
of, 71, phssical, 214, petrifaction of, 
214 working, 220 not articles of 
filth 220 

lIi/pothest% non fingo, 138, 211 

Idevhsm, 35, 81, 85 91 201, 257, rela- 
tion to positivism, 11, subjective, 14, 
193, 201 German 47, critical, 58, 
antithesis to m iteiialism, 85-86, 
movement toward, in modern physics, 
120-137, 192, transcendental, 174, 
182 propaganda for, 190, objective, 
of Hegel, 201, mcnshevizing, 203, 
two front war agiinst, 206, varieties 
of Kaiiti in, 257 
Idcilistic asscition, 112 
Idealists. 95, problem of materialists 
and 30 

Ideas, general, prejudice against, 273 
Ideologies, 258, 283, conflict of politi- 
cal, with science, 17, of new organ- 
ismic state, 126, antiscientific, 230, 
historical and contemporary, 258, ri- 
val, struggle among, 281 
Ignorabimus, 92-95 passim, 97, 99, 111, 
112 

Ignonng the facts, 264 
Imagery, geometric, 151 
Imagination, 266, free, axiomatic sys- 
tem a product of, 14, human, 58, 60, 
logical, 297 


312 



index 


Indeterminacy, principle of, 255 
Indeterminism, concept of, 255 
Indoctrination, political and religious, 
273 

Induction, law of, 270 
Inertia, law of, 53, 210, 212, 214, 235- 
236 

Inquisition, 91; condemnation of Co- 
pemican system, 208 
Integration, philosophy of, 36-37 
Intelligentsia, middle-class, 191 
Intensities, electric and magnetic field, 

59, 115 

Interlude, Kantian, 47 
Interpretations, metaphysical, 51, 164, 
300; philosophic, 159; pseudophilo- 
sophic and pseudoreligious, 231; logi- 
co-empirical, 290 
Interpreters, philosophic, 168 
Intersection, geometric concept of, 12, 

13 

Intervention, of political powers, 51 
Intolerance, spirit of, 279 
Intuition, internal, 147-148 
Intuitive theory, 150-151 
Iron curtain, between science and phi- 
losophy, 25 

Irony of embarrassment, 114 
Issues, historic, 282 
Italy, 124 

James, William, 32, 33, 40, 79, 95, 101- 
103, 111, 218, 256 
Jaumann, Gustav, 71 
jeans. Sir James, 129, 130, 133, 134, 
135, 184, 186, 188, 250 
Jesuit, 209 
Jordan, P., 184 

Journal, philosophical, of Vienna Cir- 
cle, 38 

Journal for the Whole of the Natural 
Sciences, 195 

Judgment(s), subject-predicate form of, 
104; relation of true, to reality, 105 
Jupiter, planet, 223 

Kant, Immanuel, 23, 95, 147, 180, 212, 
214, 257; theory of knowledge, 7; 
conception of geometry and me- 
chanics, 7-8; idealistic philosophy, 9, 

11; Critique of Pure Reason, 57; for- 

3T3 


mulation of law of causality, 175, 177 
Kantian interlude, 47 
Kantianism, 40, 47; fundamentalist and 
modern attitudes, 23-24; new philos- 
ophy and, 31. See also Neo-Kantian- 
ism 

Kepler, Joliannes, 63, 90 
Kirehhoff, Gustav Robert, 65 
Kleinpeter, H., 77 

Knowledge, 233; theory of, 7, 90; syn- 
thesis of, 88, 262, 271; limits of, 
9.3; cxtrasL-ientific or superscientific 
sources of, 193; explanatory and un- 
derstanding, 196; genuinely satisfy- 
ing, 196; integration of, 230, 271, 
276; prescientific, 268, 270; isolated 
branches of, 269; beyond science, 
269; useful, without truth, 271 
Kronecker, Leopold, 66, 67 

Labor, division of, 37, 260; problems, 
247 

Lacunae, in science for introduction of 
supernatural factors, 183 
Lange, F. A., 194 

Language, phenomenal, 35, 85, 87, 171; 
transition from quasi-idealistic to 
quasi-materialistic, 36; physicalistic, 
36, 161, 202; of unified science, 37, 
87; Mach’s perception, 83-84; physi- 
cal, of Neurath, 85, 87; formative 
rules of, 154; of daily affairs, 157, 
171, 179; of quantum physics, 157; 
universal, of science, 161; comple- 
mentary, 166; protocol, 171; symbol, 
of psychoanalysis, 171; intersubjec- 
tive, 202; kitchen, 243 
Laplace, Pierre Simon de, 124, 175, 
177; school of, 281 

Law(s), physical, 6, 149; statistical, 
119; deterministic, 176; concept of, 
178; metaphysical, 264 
Law of Causality and Its Limits, The 
(Frank), 176 
Leaves, withered, 64 
Leibniz, Gottfried Wilhelm von, 47, 
132, 138; Newton and, 280 
Length, measurement of, 115-117; op- 
tical illusion of, 127; relative and ab- 
solute, 154-155, 239, 246; subjectiv- 
ity of statements about, 300 



index 


‘‘Length of a lod,” 127 
Lenin, Nikolai, 10, 11, 100, 125, 190, 
191, 193, 198-199, 204, 258, 284, 
denunciation of \lachism, 17, 35, 
47 

Le Roy, E , 299 

Lessing, Gotthold Ephraim, attacks on 
Voltaire, 72 

Levi-Ci\ita, Tullio, 134, 135 
Liberalism, fight against democracy 
and, 280-281 

Life, 83, conscious intellectual, 159, 
Mtalistic conception of, 159 
Light, nature of, 93, 98, identity with 
electromagnetic radiation, 98, dif- 
fr.icted oi scattered by electrons, 
116, waselengtli of, 116, corpuscular 
theory of, 211, wase theory of, 246 
Light quanta, 116, 117 
Light rays, 107 

Limitations, absence of, in changes in 
symbol system, 113-114 
"Limitations of Natural Science” (Du 
Bois Reymond), 92 
Literature, courses in, 232 
Lobatchevski, Nikolai IvanoMch, 270 
Logic, need of pragmatic oil, 11, sym- 
bolic, 31, 39, 44 formal, role of, 103, 
of school philosophv, 103-104, Aris- 
totelian, 104, modem, 110, shift in 
the place of, 286, decline of, 1 1 twen- 
tietli century, 287, in ad\ aiice of 
tweiitietli-century science, 296, place 
of, 297 

“Logical positivism” (Blumberg and 
Feigl), 38 

Logical Syntax of Language, The 
(Cainap), 86 

Logico-empirical interpretation, 290 
Logistics, a formalistic game, 205 
Longing, unfulfilled, 159 
Loops, traced by planets, 217 
Lorentz, H A , 59 
Lutheran tlieologian, 219 
Lyric poetry, 112 


Mach, Ernst, 32, 46, 110, 140, 141, 250, 
252, 256, 298, conception of explana- 
tion, G, antimetaphysical tendencies, 
7, cnbcism of New ton’s mechanics, 8 


theory oi general piinciplcs of sci- 
ence, 11, integration with Fomcaiu 
and Einstein, 11-15, 33, 44, Soviet 
attack on, 17, 35, 47, philosophy of, 
17-18, 35, 37, 72-89, 99-100, 102- 
103, 121, 224, 225, 265, 291, 297, 
fascination of teacliings, 61, position 
of, in contemporary intellectual life, 
61-67, value of doctrines, 67-72, fol- 
lowers of, 192, 198, positivistic re- 
quirement, 294 

Machism, 66, 198 denounced by Lenin, 
17, 35, 47, niclaphvsically conceived, 
194 

Magazines, populai science, 230 
Magnetic field mtcnsitiis, 59 
Magnetism, 58 

Maimonedes, Mo‘cs, 221-222 
Man, primiti' c, world picture of, 80 
Man-in-the-strect, 156, 268, 300 
Mankind, presiientific cvpericiicc of, 
269 

Marburg, nco-Kaiitian school of, 194 
Mantain, Jacques, 24, 175, 297-298 
Mus, planet, 229 

Marx, Karl, 125, 202, 257, 266, 284 
Maixism, 190, 272 
Masar\ k, Thomas G , 45 
Mass, real and apparent, 160, meta- 
physical principle of, 161, electro- 
magnetic, 161, concept of, 255 
Matciial idiom, 152 
M itcrialism, 74, 78, 81, 85, 190, 257, 
258, 281, methodical, 35, dialectical, 
47 88, 125, 160, 198, 203, 204, 257, 
258, 271, 273, 273, 283, struggle 
against, 80, antithesis to idealism, 
85-86, in Germany and France, 125, 
dangers for, 191, strictly taboo, 193, 
metaphysically conceived, 194, mech- 
anistic, 201, 257, political systems 
based on, 263, refutation of, 280 
Materuihsm and Empinocnhcam 
(Lenin), 11, 190, 193 
Materialists, problem of idealists and, 
30, false philosophy of, 132, appre- 
hensions of, 186, scientifically 
mmded, 194 

Mathematical Pnnctplei of Natural 
Phdosophy (Newton), 212 
Mathematically true, 214 


314 



index 


Mathemabcians, meeting of physicists 
and, Prague, 39-41, conception of 
universe, 124. 

Mathematics, foundations of, 31, 103, 
spiiitual element, 123, 129, analytic 
character of theorems, 134, pure, 
propositions of, 134, concepts, 135, 
fetish, 192, replacement of mechan- 
ics, 192, relationship with physics, 

194, 234, traditional teaching of, 253, 
264, educational values intrinsic m, 

271, creative principle in, 297 
Matnces, Heisenberg’s, 292 

Matter, 83, Neurath’s avoidance of 
word, 35, and energy, 66, an auxil- 
iary concept, 74, nature of, 93, 99, 
electromagnetic picture of, 137, only 
an illusion, 160, conception of, 
changes hy electromagnetic theones, 
190, disappears, 190 
Maxwell, James Clerk, 63, 70, 139-142, 
150, 239 

Meaning, conception of, 32, 33, crite 
non of, 44, 295, theories of, 44-43, 
291, 292, definite, 154, deeper, 174 
operational, 224, 225, 244, physical, 
235 

Meaninglessness, Bridgman’s concept of, 
44 

Measurement, of space and time in 
Einstein’s theorv, 19, traditional sys- 
tems of, 22, “correct,” 96, methods 
of, 115-117, refinement of tech- 
nique, 116, 120, role of obseivei, 
127-128, physical, 243 
Measuring tod, 116 
Mech mica! engineering, 237 
Mccli lines, Newtonian, 8, 9, 23, 73, 114, 
136, 160, 213, 224, 239, 234, 270, 

272, theological, animistic, and mys- 
tical points of view in, 72, wave, 118, 
131, classical .ind relativistic, 153, 
155, replacement by mathematics 
192, traditional college, 239, founda- 
tions of, 240, 241, 245, laws of, 244, 
Einsteinian, 272 See also Quantum 
mechanics 

Mechanics (Mach), 71-72, 82 
Mechanism, question of vitalism or, 59, 
60, breakdown of, into positivism, 

195, superiority to positivistic em- 


piricism, 196, Euclidean science, 196, 
two-front war against, 206 
Mechanistic scheme, 257 
Mechanistic science See Science, niech- 
anisbc 

Mediterranean races, peculiarities of, 
148 

Medium, vibrations of, 120 
Mental constitution, 151 
Metaphysical systems, crude, 265 
Metaphysician, rabonalisbc, 302 
Metaphysics, 193, 238, 295, relabons to 
science and religion, 15, Duhem’s dis- 
bnction between physics and, 16, 
Arivtotehan, 16, 24, Thomisbc, 16, 
26, 298, traditional, 23, 34, 36, Kant- 
ian, 24, 26, 47, statements of, as social 
phenomena, 34, elimination of, 34- 
35, 82-85, 89, Schhek’s view of, 41- 
42, justification of, 42, idealistic, 86, 
88, materialistic, 86, 159, impor- 
tance of, to science, 87, spiritualisbc, 
132, 159, critical attitude towaid, 
181, scientific, 187, twentieth-cen- 
tury tendency toward, 192, German 
inclination toward, 212, sbaight, 256, 
organismic, 258, dcclme of, 290, con- 
cepbon of, 290, interpretation of 
principles m terms of language of 
everyday life, 290, heuristic value, 
297, 298, characteristic of, 299, lan- 
guage used in, 301 

“Metaphysics of Modem Physics” 
(Wen^l), 189 
Methods, experimental, 42 
Michelson’s experiment, 254 
Middle Ages, 124, 186, 286, animistic 
science of, 122-123, language of, 211 
Milky Wav, 226, 237 
Mill, John Stuart, 9, 12 
Mind, nature of human, 7, ohjeebve ex- 
peiience and, 9, fiee creabons of, m 
Poincare’s theory of science, 12, Neu- 
ratli’s avoidance of word, 35, eman- 
cipation of, 69, gulf between body 
and, 122, problem of explaining, 122, 
creabvity of human, 278 
Mismterpretabons, of physical theones, 
46-47, philosophic, 160, metaphysi- 
cal, 164, 165 
Mitm, M , 191 


315 



index 


Models, mechanical, 150, 152 
Momentum, conservation of, 164. 
Morris, Charles H , 48, 49, 290, 300 
Motion, Newton’s laws of, 8, 9, 145, 
235, positivistic doctrine of space and, 
68, absolute velocities ot, 117, of 
bodies of average size with moder- 
ate velocities, 131, sunple 208, triple, 
208, indestructibility ot, 212, abso- 
lute, expenditure ot divine energv, 
224, real and apparent, 238, essence 
of, 270 

Nagel, E , 48 

Narrowness, mechanistic, 199 
Nationality, 150 

Natural laws, fuimulation of, 14, mali- 
cious, 182 

Natural Science on the Path to Re- 
ligion (Baviiik), 123 
Nature, simplicity of, 11, poitrayil of, 
57, fictitious, 5S, empirical, 59, scho- 
lastic conception of, 90 suirender of 
scientific conception of, 97, scientific 
conception of, 110, insidious hws of, 
117, spintualistic conception of, 124 

137, organismic conception of, 124, 
processes of, mechanical and mathe- 
matical bases, 130, new physical con- 
ception of, 136, 159, existence of 
sunple laws, 176, mathemitical con- 
ception of, 188, idealistic philosophv 
of, 188, mechanistic conception ot, 
287 

Nazism, 248, 263 
Negabon of the negation, 202 
Negativism, positivism not a, 68 
Neo-Anstotehamsm, 23-25 
Neo-Kantiamsm, 23-25, 43 
Neo-Tliomism, 23-25, 270 
Neo-Thomists, French, 48-49 
Neurath, Otto, 1, 2, 31, 34, 38, 85, 86, 
161, 166, 171, 202, index of pri- 
hibited words, 35, unification of sci- 
ence, 36-38, 39, 47 
Newspapers, 230 

Newton, Sir Isaac, 63, 73, 132, 135, 
159, 186, 210, 211, 223, 286, prin- 
ciples of mechanics, 23, physics of, 
115, 138, antimatenahsbc tendency, 

138, equations, 242, contemporaries. 


254, popularization of, 263, issues 
between Deseartes and, 279, Leibnu 
and, 280 

Newtonian, Newton not a faithful, 211 
Newtonian laws, 8, 9, 145, 156, 213, 
214, 223, 225, 235, 237, presentation 
of, 254, results of evpenence of every- 
diy life, 289 

Newtonians, conflict with Einstein, 268 
Nietzsche, Friedrich, 9, 18, 75, 76-78 
Nineteenth century, 136, 212, 216, 220, 
224 261, 280, 286, ph)sical research, 
149 

Nummahsm, 252 
Noninteivention, pohej of, 269 
Nordic r ice, 248, characteristics of. 1 13 
Northrop F S C , 42-43, 244, 258 
Numheis, iriitiunal, 66, 108, 109, 110, 
nitiire of complex, 100, i itiuiiil, con- 
vergent sequent e of, 109 
Nurimberg, 219, 222 
Nurse, 265 

Objetl(s), agreement bitwtcii thought 
and 98, psyehologieal. 111, physic il, 
170 

Obscurantism, souice of, 267 
Obseivabon, 242, nme sensoi), 146, 
objective, of human actions, 166 
agreement of consequences with, 210, 
sense, 218 

Observ ational teims, v ..gueness of, 1 3 

Observer, role of, 127-128, 254 

Observing subject, 126 

Occam, William of, 284 

Oec im’s razor, 252 

Ontology, 41 

Oper itionism, Biidgman’s, 258 
Optical illusion, 217, of length, 127 
Optics, statements ot, 245 
Oiganicism, 195, glorification of mech- 
anistic physics by advocates of, 196 
Organism, physical disturbances of, 168 
Osi inder, Andreas, 219-220, 222 
Overspeciah/ation, superficiality and, 
259 

Oxford, 294 

Faiallels, axiom of, 134 
Paralogisms of rationalism, 48 
Fans, 49 


316 



Index 


Particle ( 9 ), as term m atomic physics, 
170, small, with large velocities, 119 
Paul III, Pope, 222 
Peirce, C S , 32, 48, 256, 300 
Peiception terms, 84 
Perceptions, sense, 68, comple\es ot, 80 
Pcifection, spiritual, 161 
Periodicals, 230 
Persecution, political, 50 
PetnfaePons, 214, of Newton’s physics, 
215 

Phenomena, true cognition of electrical, 
106, understanding of, 138, electro- 
magnetic, 139, as energy transforma- 
tions, 158 
Phenomenalism, 70 

Philosophers, 253, professional, 76, task 
of, 114, idealistic, 147, Geimaii pro- 
fessional, 187, scholastic, 209, diser 
gence in attitude of physicists and, 
210, tvsentieth-eeiitury, 210 
Philosophic schools, reactionary, 191 
Philosophical journal ot Vienna Circle, 
38 

Philosophically false, 214, 215, 218, 226 
Philosophically true, 219, 221, 225 
Philosophy, 255, 276, Catholic, 2, sci- 
ence and, 4-5, 25 121, 207, 262, 265, 

266, 269, 273, 275, haditum.d, 14-15, 
23, 25, 28-29, 32, 43, 81, 82. 85. 88, 
idealisbc, 18, 33, 80, 85, 160, 162, 
165, 300, fundamentalist and mod- 
el nist, 23, 25. Ceiman, 34, 47, ana- 
lytic and speculative, 36, of integra- 
tion, 36-37, Comte's definition of, 37, 
dislike of term by Vienna group, 38, 
new, of Vienna group, 38-39, turn in, 
41-12, Aristotelian, 73, 211, 219, 220, 
mathematical principles of, 132 
atomistic, 142, orthodox, 149, contro- 
\ersial questions of, 158, Kantian, 
180 official, 193, orgamsmic, re- 
garded as Cermm, 195, empirical, 
212, 232, 242, departments of, 262, 

267, relation to physics, 263, popular, 
265, chance, 265, 287, 268, 276, ob- 
solete m socifil and political life, 266, 

268, special field of, 267, 268, 273, 
275, vuerage instruction in, 267, as 
integration to human knowledge, 270, 
traditional teaching of. 271, mter- 


acbon with science and rehgion, 279 
See also School philosophy, Soviet 
Union, philosophy of 
Philosophy of ‘As If (Vaihinger), 43 
Physical law, Mach’s theory of, 6, new 
types of, 149 

Physical terms, meaning of, 44 
Pliysical theories, conception of, 19, 
misinterpretations of, 46-47, fossiliza- 
tion of, 115, modern, standpoint of 
school philosophy, 119, philosophical 
mterpietation of, 160 
Physicahsm, 35, 36, 85, 166 
Physicalists, 160 

Physicists, meeting of mathematicians 
and, Prague, 39—41, poor training of, 
in philosophy, 46, reactions of, to 
modern theories, 97, theoretical, work 
of, 1 13, stimulus to, 146, progressis e, 
151, di\ ergeiice in atbtude of philoso- 
phers and, 210, twentieth-centiuy, 
226, named, 230, experimental, gul- 
libility of, 231, failure of learned, 
264, lack ol philosophy of physics 
among, 266, average, 274 
Physics, 76, 122, 140, anthropomorphic 
conception of, 159, Aristotelian, 73, 
208, 209, 215, 219, 222, 223, 252, 
as church, 78, atomic 240, comple- 
ment'’rity conception, 179, no longer 
deterministic, 177, nuclear and, 238, 
world of, 157, classical 114, 115, 120, 
127, as extinct organism, 280, crisis 
ot, 280, epistemology of, 119, college, 
elementary courses, 284, common- 
sense, 241, concepts of, 37, crisis of, 
2, 136, 215, Duhem's distinction be- 
tween metaphssics and, 16, educa- 
tional values intrmsic in, 271, Ein- 
stein’s non-Aristotelian, 74, estab- 
lished prmciples of, 223, experimen- 
tal, 44, graduates in, 264, influence 
of philosophical creeds, 23, laws of 
59, gaps in, 183, logico-empirical, 
253, Mach’s news of, 6-8, mate- 
rialistic end of, 137, idealistic ele- 
ment inside, 104, mechanisUc asser- 
tion of ontologic value, 2, concepts of, 
18, decline of, 2-3, 252, 280, orig- 
inal sin, 123, overestimation of, m 
behalf of orgamsmic science, 195, 


317 



index 


196, penod of, 251-232, medieval, 
139, 221, Newtonian, 132, 215, 280, 
Hint teenth-Lcntiirv 139, 282, Ein- 
stein’s theory and, 20, nonmechaincal, 
mathematical, 131, orgmismic dw- 
mtegration of, 252, return to, 282, 
place of, in human thought, 5-6, 
poMtiMstic 140, as \ariety of subjec- 
ti\c idealism, 193, relationships with 
mathematics, 194, 234, philosophj, 
263, psychology, 69, sociology, 272, 
spiritualistic, 128, 159 teaching of, 
230 if , 264, 267, textbooks, 264, 278 
traditional, 232, tw entieth-century 
114, 120, 149, 215, 217, 231, 252, 
253, 263, 282, 287, histoiy ot, 292, 
influence on human affairs, 264, inter- 
pretition of, 50, 231, 296, language 
of, 299, new logical techniques, 254, 
rebirth of idealistic, 123-137, 140, 
186, understanding of, 96, undti- 
standing of, 233-234 S e aho Sci 
ence 

Phisiologi, 83 
Pick, George, 72 
Pictorial representation, 147 
Picturil theory, 139, 147, 148 
Picturization, 139, 147 
Planck, Max, 62-70 pfissim, 76, 166, 
250 

Planck constant ii, 102, 107 
Planetary orbits, perturbations of the, 
145 

Poetry, as personal experience, 42, lyric, 
112 

Poincare, Henri, 17, 26, 30, 46, 48, 53, 
54, 69, 100, 250, 256, 299, new posi- 
tivism, 8-12, 20-21, structural sys 
tern m physical theory, 13-14, inte- 
gration with Mach and Einstein, 14- 
15, 33, 44 

Point, geometric concept of, 12, 13 
Point experiences, 119 
Pointer readings, 107, 128, 243 
Poland, 18, 26 

Political ideologies, conflict with sci- 
ence, 17 

Political problems, understanding of, 
232 

Politicians, global, 263 

Politics, 271, 277, factors of, 255, 


slogans m, 266, syntliesis with sci- 
ence, ethics, and religion, 271 
Popper, Josef, 72 
Popularizations, superficial, 268 
Position! s), initial velocity and, of elec- 
tron, 118, indefinite, 103, new syn- 
tax for word, 179 

Positivism, 223, 251, breakdown of 
mechanism into, 195, cause of, 
against metaphysics, 23, Central Eu- 
ropean, 48, 49, conneebon witli uni- 
ficabon of science, 37, dislike of term 
b\ Vienna group, 38 evolution of, 38, 
logical, 39, 42, 43, 50, 256, mistake 
of, 196 new 24, 26, 34, 35, 47, 
Einstein’s theory and, 18, 20-21, 28, 
French advocates, 48, Hilbert’s con- 
tiibubon to, 13, mtegrabon of philos- 
ophy of Mach, 14, mergence with 
Thomistic metaphysics, 16, of Poin- 
care, 9-12, 13, spirit of, 5, 6, nine- 
teenth cenbiry, 178, not a negativism, 
68, relation to idealism, 11, straight, 
256, Study’s view of, 65, the rape uti- 
c il, of Wittgenstein, 32 
Positivists, 252 

Pragmitism, 33, 102, 111, 223, 256, 
Amencin, 40, 48, 203, Central Euio 
pc in ijositivism and, 48, truth-con- 
cept of, 112 

Piagmatim (Janies), 95, 101 
Prague, 47, 49, 99, meeting of piiysicists 
and mathematicians, 39-41, Univer- 
sity of, 31, 45, 47, 99 
Preachers, 263 

dedicate, subject and, in school logic, 
104 

Prediction, of future processes, 177 
Predilections, religious and political, 256 
Pre-Gahlcan penod, 287 
Prejudices, popular, 9, metaphysic il, 
173, 240, obsolete, 264 
Prescienhfic stuff, 269 
Presentations, popular, 164 
Presuppositions, 270 
Principles, tautological tiansformations 
of, 110, Newton’s, 122, metaphysical 
209, plnlosophic, 214, 215, checked 
by extrascientific metliods, 251 
Probability, mathematical, not a physi- 
cal reality, 187 


318 



Probability concept, spiritual factor, 158 
Problems, metaphysical, cognition in so- 
lution of, 30, verbal, of traditional 
philosophy, 32, soluble m prmciple, 
111, pseudo, 162 

Processes, prediction of future, 177 
Professionid training, 230 
Propaganda groups, 264 
Properties, perceptible and impercep- 
tible, of bodies, 56 

Proof, mathematical, and experimental 
confirmabon, 234 

Froposibons, tautological system of, 24, 
metaphysical, 84, 88, meaningless 
metaphysical, 165, isolated, 173 
Protocol language, 171 
Pseudo-idea, 212 

Psychic factors, in interpretataon of 
physics, 126 

Psvchoanalysis, symbol language of, 171 
Psychologic fact, 213 
Psychological elements, in me isure- 
ment, 127, 128 

Psychology, 83, 84, 271, concepts of, 
37, connection of, with physics, 69, 
individual, ISO, understanding of, in 
terms of ph)sics, 165, empincal, 170, 
application of complementaiy prin- 
ciple to, 180 

“Psychology of Metaphysics” (Nietz- 
sche), 77 

Ptolemaic system, 73. 143, 220, 221 
Ptolemy, 208, 218 
Punishment, 167 

Quacks, 240 
Quanta, 67, 230 
Quantities, observable, 107 
Quantization laws, 240 
Quantum hypothesis, 116 
Quantum mechanics, 114, 115, 120, 125, 
128, 131, 153, 163, 292, principles of, 
96, philosophic misinterpretahons, 
162, syntactic form in, 180, relation 
to ethics 184 

Quantum of action, real existence of, 
108, 109, 110 

Quantum physics, laws of, 149, lan- 
guige of, 157 

Qii intum theoiy, 40, 126, 127, 130, 140, 
148, 210, 215, 240, 252, 255, 267, 


268, creation of, 123, mdetennin- 
ishc mterpretahon of, 128, research 
on, 151, analogy of free will to, 
166, posibvisbc concepbon of, 178, 
metaphysical interpretabon of, 182, 
253, misuse of, 184, irrabona] as- 
pect, 246 
Qume, W. V, 48 

Race, 150, 195 

Race relabons, problem of, 247 
Radar, 262 

Radiabon, 08, 107, 292 
Rationalism, illusion of, 5, new and ba- 
dibonal, 9, paralogisms of, 48, 
French, 195 
Rationalists, 74 

Reactions, of physicists to modern theo- 
ries, 97 

Real, use of, is word, 35, 162 
Realism, 17, 94 
Realist, 95 

Reality, 62, objective, 14-15, 29, physi- 
cal, 18, 178, 215, 221, ontological, 18, 
electric forces as elements of, 27, 
Neurath s avoidance of word, 35, 
symbolic logic and its application to, 

39, hidden iii a nutshell, 98 re- 
lation of bue judgment to, 105, metv- 
physical, 108, iionmechanical, 129, 
concrete, 136, dual, 165 

Re.'soii(s), pure, 212, 214, psychologic, 
248, extrascientiEc, 248, 253, 256, 
257, marriage of fact and, 286, dt- 
chne of, in twenbetli century, 287 
Riasoiniig, experience and, in geoni 
cUv, 21-23 speculative, 218 
Recurrence of a state of a system, 14 
Reference, system of, 227, 235, 254 
Heichenbach, H , 26, 30-35 passim, 

40, 41, 44, 47, 48, 112, 115, 245 
Relabon(s}, relation of, 10, of cooidina- 

bon, 44, bebveen two things, 104, 
analogical, 290, umvocal, 290 
Relativism, Mach accused of, 17, of 
modern science, 52 
Relativity, theory of, 18-21, 26, 40, 46, 
68, 73-74, 96, 97, 114^128 passim, 
131, 135, 140, 148, 150, 153, 155, 
162, 191, 210, 215, 217, 225, 226. 
238, 245, 252, 253, 266-268, 27< 



index 


275, misunderstanding of, 127, popu- 
lar literature on, 127, laws of, 149, 
teachmg of, 238-239, 240, restiicted 
theory of, 254, theory of meaning 
and, 291 

"Relativity— a richer truth,” 52 
Relativity boom, 229, 262 
Religion, 24, 193, 272, 277, relations 
to science and metaphysics, 15, re- 
lations with science, 17, 123, 230, 
259, scientific world conception and, 
40, mi\ing of, into science, 224, re- 
lations between science, goienimenl, 
and, 232. factors of, 255, ssnthesis 
of science, ethics, politics, ind, 271, 
interaction with science and philoso- 
phy, 279 

Representation, pictural, 139 
Research, spirit of, 63, e\act scientific, 
restricted to dead matter, 196 
Responsibilitv, moral, 183 
Restriction of expectation, 65 
Ke\, Abel, 2, 5, 9, 14, 46. 69 
Ricci, C G . 134, 135 
Riddle(s), eternal, 92, insoluble, 93 
Riemaiin, G F B , 130, 134, 135, 136 
Roman Church, 272, conflict with 
topeniie in s)stem, 16, 47, 218, 226, 
232, 268, cardinals of, 258 
Rougier, L , 48, 49, 202 
Russell, Bertrand, 33, 46, 103, 104, 110, 
133 

Russia, overthrow of Czarist regime, 26, 
216 See aho Sonet Union 
Ruver, R , 123 

Sanity, intellectual, 6 
Scale diMSioiis, coincidences of point- 
eis and, 107 

Scheme, oversmiphfied, 269 
Schhclk, Moritz, 26, 38, 48, 83, 105-106, 
113, 183, criterion of truth, 28-29, 
cognition as establishment of cor- 
respondence, 29-30, cooperation w ith 
Viennese group, 32-36, “The Turn 
in Philosophy,” 41—42, “Space and 
Time,” 43, ass.'issinatioii of, 49, vi- 
cious philosophi, 49 
Scholasticism See School philosophy 
School phi’asophy, 8, 78, 94-121, 163, 
189, scientific world conception and. 


40, tradibons of, 97, new fashionable 
form of, 100, truth concept of, 101, 
105, logic of, 103-104, time concept 
of, 114, conceptions of, 117, disin- 
tegration of, 174, 175, 181, social 
function, 200 

Schioder, 103 

Sclirodmger, Edwin, 118, 179, 292, 295, 
296 

Science(s), absence of boundary points, 
111, all-round understanding of, 248, 
animistic of Middle Ages, 122, re- 
turn to, 194-195, as dynamic living 
being, 280, as economical represen- 
tation of cxpeiieaccs, 84, as game 
with empty sjmbols, 199, bourgeois 
conception of, 205, cimpaign against 
spirit of, 246, collapse of inneteeiith- 
century beliefs, 4, conception of, as 
useful technique, 3, crisis of, 215, 
critics of, 260, cross connections be- 
tween branches of, 273, decadence 
of, 125, disontologization, 175, eman- 
cipatory lole of, 5, eiolution of, 290, 
foundations of, 284, Geiman, 47, 195, 
gospel of, 268, heritage of ancient 
state of, 238, histoiv of 149, 278, 
turning point m, 280, idealistic in- 
terpretition of, 195 inherent educa- 
tional value, 247, integration of the. 
269, 273, 276, intrinsic human values, 
261, lacunae for introduction of su- 
pcriiatuial factois, 183, limit itions of 
sbitcnicnts of, 15, logic of, 220, 23S, 
mechanistic failure of, 2-5, M icli’s 
debunking of, 18, medieval, oiei- 
tlirow of, 223 met iphysieal intei- 
pretitioiis of 11, 23, 51, 87, 88, 256 
289, 290, org inism of, 62 philosophv 
of 45, 50, 220, 253, 273, 277, 27S, 
284, cultur il background and, 42, im- 
portance of, for political creed, 258, 
te idling of, 51, 249-250, 258, train- 
ing in, 274, physical 208, public In- 
tel est 111 , 262, place of reason in, 289, 
position of. 111 cultural life, 256, posi- 
tivistic 166, 175, fight against, 195, 
piinciples of application of common 
sense to, 300, 301, Mach’s and Pom- 
car4’s theories, 11-12, pseudo-posi- 
tiv ist conception of, 87, pure, 57, 97, 


320 



index 


relabonships with- humambes, 258, 
261, 281, 284, philosophy, 4-5, 25, 
121, 207, 262, 266, 269, 273, 275, 
279, rehgion, 15, 17, 123, 224, 230, 
232, 259, return to pre-Galilean, 195, 
sociology of, 200, special, binding 
materim between, 269, symbobc 
structure of, 105, 279, synthesis with 
ethics, pohbcs, and rebgion, 271, 
teaching of 229, average ihstrucbon, 
267, conbibubon of, 278, conven- 
tional, 230, courses, 232, gaps m ba- 
ditional, 245, 250, 253, improvement 
of, 247, pedagogic effort, 217, short- 
comings, 231, 249, theorebcal, fun- 
damental questions of, 58, badibonal, 
84, 88, twentieth-century 25-26, 40, 
45, 289, relabsism of, 52, semanbe 
significance of, 247, unity of, 35, 36- 
33, 79-89, 174, 193, 197, 202, uiii- 
\ersal language of, 83, 87, 161, un- 
obsersable ilcmtnts in, 39, Victoiian, 
184 

Science and God (Basink.), 187 
Science and llijputhcsis (Poinciie), 53 
Science columns, newspaper, 230 
Science, Pliilosopli) ind Hcligioii, Con- 
feience of, 52 

Science student, average, 261 
Scientific method, antiscicntific tenden 
tits, 2 

Scientific piiisuits, cans 1 connection 
witli otlier social processes, 200 
Scientific world conception, of Vicinii 
group, 3S-1S, 68, 92, 94, 108, 112 
114 

Scienbsin, illusion of, 5 
Scientists, educition of future, 258 it- 
titude of !e iding 261 contiibution to 
political life, 261, nitural, uncLr 
domination of philosophy, 266 
Self-observation, 166 
Semantics, 214, apprnacli, 45 300, m il 
adjustment, 241, rules, 244, 246, 291, 
analysis, 245, 249, training in, 249 
Semites, 148 
Sense d ita, 14, 56 
Sense perceptions, 68 
Senses, evidence of, 145 
Sermons, 263 
1789, ideas of, 124 


Seventeenth century, 290 
Shapley, Harlow, 52, 226 
Smiphcity, concept of, 176, require- 
ments of, 296 

Simultaneity, m absolute and relative 
sense, 155 

Simultaneously, use of, as term, 155 
Situabons, complementary, 166, 167 
Sixteenth century, 207, 225 
Skepbcism, 14, 120, threatemng, 68, 
overthrow of enlightenment by, 74- 
75 

Slogans, debunking empty, 264, gen- 
eial, in pohbcs, 266, philosophic, 282 
Smuts, General Jan Christiaan, 124 
Social phenomena, metaphysical state- 
ments as, 34 

Social problems, understanding of, 232 
Societ}, human, crisis m, 136 
Socio ps) chologic approach 250 
Sociologists, 253 

Sociology, piisitiv istic, eudaemonistic 
etliics and, 39, of science, 200, dis- 
closes new laws of matter, 202, laws 
of, 271, synthesis of physics and, 272 
Sommcrfeld, A , 164 
Soul, collective, 83, science and the, 
125 

Soviet Union, 271 establislunent of, 17, 
IS, Mich atticked by, 17, 35, 47, 
philosoph) of, 35, 47, 88, 125-126, 
160, 198-206, 257, anb-Machistic 
school, 100, authors quoted, 186, 
187-189, 190-192, materialistic lit- 
er ihiie, 192 

Spice, 112, 26S, measurement of, 19, 
22, lour dimeiisiniial, 28, 158, and 
mohon, positivistic docbine of, 68, 
real, 114, fianie of physical phe- 
nomena, 115, 120, Riemanman, 130, 
134, 135, 136, form of experience, 
180, Euclidein and non-Euchdean, 
196, 233, discussions of, 207, empty, 
211, absolute, 223, 225, 238, 239, 
nature of, 229, 253 
“Space and Time” (Schhek), 43 
Space-tmie continuum, four-dimen- 
sional, 152 

Specialisation, cersus ^neral educa- 
bon, 261, belief of saenbsts in, ‘273 
Speech, material mode of, 162 


321 



Index 


Speeches, on festive occasions, 173 
Spencer, Heibert, 212, 214, 257, 302 
Spintiuilisni, 81, 124, 126 
Spontaneity, of action, 162 
Spoon feeding, 281 
Stallo, 252, 256 
St irs, 225 

State of affairs, definition of the, 176 
State of a system, 56 
Statements, purely logical, 252, intro- 
ducing new words, 242, ohserca- 
tional, 243, 244, tautological, 243 
common-sense, 300 
Statistical element, 118 
Stohr, Adolf, 58, 71 
Straight line, geometiic coiuipt of, 12, 
13 

Strauss, M , 179 

Structural s)stem, 44, in ph\ steal the- 
ory, 12-15, in Einstein’s lliioiv, 20 
Euclidean and non-EticlidL in 22 
Student accrage, of pin sics 203 
Studs, E , 62, 65, 66, 6S, 76 
Subconscious, 248 

Subject, predicate and in school logu 
104 obsers mg, 170 
Subjectn ism, Mach icci std of, 17 
Subject-predicate foim ol judgments 
104 

Substance, material notion ol, 187 
Substantiation, intersubjcilni in imas- 
uremeiit, 127, 128 
Sun, 226, 227 
Sunday sermon, 41 

Suptificiality, oierspeciali/atioii and, 
259, 273 

Supernatural, lacunae in science loi in 
troduction of, 183 

Svmbols, abstract, 13, in Einstein’s the 
oi\, 20, sjstem of, 29-30, 105, 107, 
113-114, 120, 243, cognition as sys- 
tem ol, 42, experienees and, 106, 
118-119, operational meaning of, 
248, stiucture of, 278, \alue of, 282, 
priiieiples couched in, 283, liave own 
hie, 283, choice of, 283 
Ssiitaetical differences, between classi- 
cal and relativistic mechanics, 155 
Syntax, relativistic, formative rules of, 
>56, logical^ of Carnap, 162, rules of, 
163 


Syndiesis, of human knowledge, 271, 
of physics and sociology, 272 
System, state of a, 56, 120 


Tam, llippolyte, 9 
Talk, idle philosophic, 264 
Tautologies, 110 

Teachers, gap in training of, 239-240, 
physics, 264 

Teaching See Science, teaching of 
Technical material stuffing with, 281 
Teleological element, in quantum me- 
chanics, 128 
Television, 262 
Tendencies, antiscientific, 2 
Teiider-mmded, 218 
Tenuis, ordinary rules of, 156 
Terminology, question of, 59, Kantian, 
180 

Terms, vagueness of obsiivalion il, 13 
Tencstnal bodies, 221 
lest body, 117 

‘ Ttstabilitv and Meaiimg” (t 'map), 

86 

Textbooks, in Soviet Union, 191, 199, 
presentation of Copeinican conflict, 
231-232, jahvsics, 234, 235, 264, 278, 
non-Euclidean geonietrv, 234, law ol 
nieitia, 236, tradition il, 252, cosniol- 
og\, 256, di ilcctic d m itciialisin, 272, 
current, 280 
Theologians, 253 
rheology Set Religion 
Theories, scientific, metaphysical inter- 
prctitions ol, 11, twentieth century, 
16, mi t iphv sical, 46 truth of, 107, 
matcii ilistic social 136, distiiicUoii 
between meclnnstic and matheniati- 
crl, 139, distinction between iiitui 
tive and abstract, 146, 149, artificial, 
153, prescientific, as philosophic 
world picture, 160, transition from 
mechanisbc to mathematical, 194, 
pniduction of, by free imagination, 
295 

Thing(s), Neurath’s avoidance of word, 
35, property of a single, 104, relation 
between two, 104, loncept of, 178 
Thing m itself, 162 
Thing language, 37, 152 


322 



index 


Thinking, progressive, 48, didlccticiil, 
203, rational, crisis of, 280 
Thomism, 256, 258, 271, 273, 283, 
Duhcm’s advocacy of, 16, chief asset 
of, 272 See also Neo-Thomism 
1 homistie Congress, Rome, 1936, 175 
Thomson, J. J , 158, 160, 161 
Thought, economy of, 65, agreement 
between object and, 98, 102, 105, 
disintegration of rational, 121, pure, 
umcerse consisting of, 187, dialecti- 
cal laws of, as laws for matter, 202, 
scientific, rising influence of, 260 
Time, 112, 268, measurement of, 19, 22, 

114, true, real, 114, framework of, 

115, 120, form of experience, 180, 
discussions of 207, nature of space 
and, 229, absolute, 239, lelatn ity of, 
255 

Time scale, Einstein, 113 
Tolerance, 281 
Totalitarian groups, 34 
Tough-minded, 218 
fiactatus Logtco-Philosophitus (Witt- 
genstein), 31-32 
Tradition, standard, 271 
Tiansformation, logical, 134 
Transition from quantity to quality, 202 
Trends, American philosophical, 48, re- 
ligious, social and political, 248 
liiangle, 243, physical, 134, essential 
pioperties of, 147 

Truth, 295, Schlick’s criteiion of, 28-29, 
Hcichcnbach’s ciiterion of, 31, philo- 
sophical 91, piactic.d and theoretical 
conceptions of, 102, an invention, 
102, criteria of, 102, 108, 219, 222, 
223, 251, absurdity of pragmatic con- 
cept of, 104, eternal, 114, 115, con- 
cictc, 200, 203, scientific and philo- 
sophic, 209, 211, 220, mathematical, 
211, two conceptions of, 219, dis- 
crepancy between mathematical and 
philosopinc, 222, relativity of, 230, 
scientific and philosophic, 251 
“Turn in Philosophy, The” (Schlick), 
41 

Twelfth century, 221 
Twentieth century, 224-227, 256, 261, 
286, physical theoiies of, 123-137, 
physictd research, 149, antiscientific 


ideology, 230, obscure domain of tlie- 
oiics, 241 

Uncutainty, piinciple of, 168, 240, 275 
Unconscious, unfulfilled longmg m, 159 
Under the Banner of Marxism, 191 
Understandmg, conception of, 32, em- 
pithic, 197, casual, 213 
Undeistandmg Science (Conant), 50 
Unified field theories, 295 
Unified scheme, 301 
United States, 33, 44, 48, 50 
Unity, lost, of nature, 159 
Unity of science, 35, 36-38, 79-89, 174, 
193, 197, 202 

Unity of Scieiiie, Congress fin the, 49 
Universe, “truth” about, 4, mathema- 
tician’s conception of, 124, spiiitual- 
istic conception of, 160, consisting of 
pure tlioiight, 187, true picture of, 
222, central stage of, 227, master of, 
227, democratic order of, 227, intel- 
ligible by analogy, 298 
Universities, German, 34 
Utopianism, 65 

Vagueness, 243 
Vaminger, H , 43-44 
Valentinov , N , 190 
Validity, eternal, 215, 280 
Value(s), emotional, 59, judgments of, 
77, observed and calculated, 107, 
exact, 115, 116, educational, 284 
Value of Science, The (Pomcare), 53 
Variables, fictitious state, 60 
Velocity, initial position and, of elec- 
tron, 118, with respect to the ether, 
154, 155, new syntax for, 179 
Venfication, operabonal, 43, direct ex- 
perimental, 140 
Vibrations, of a medium, 120 
Vienna, City of, 31, 48 
Vienna, University of, 1, 2, 32, 47, 49 
Vienna Circle, 1-2, 31, 33, 34, 49, 79, 
85, 86, 133, 160, 172, 203, coopera- 
tion of Schlick and Carnap, 32-36, 
elimination of metaphysics, 34-35, 
unification of science, 36-38, at- 
tacked as anbphilosophic, 175 See 
also Scientific world ooncepbon , 
Viennese waltz, 38 


323 



Index 


Viewpoint, prcpositivistic and prescien- 
tific, 42 

Visualizable theory, 146 

Vitalism, question oi mechanism or, 60; 

spiritualistic, 168 
Voltaire, Frangois Arouet, 72, 75 
Vouillcmin, General C. E., 49 


War, causes of, 260 
Wave functions, 118, 202 
Wave mechanics, 118, 131 
Weimar Republic, 26 
Wells, H. G., 228-229, 232, 341 
Weltanschauung, 38 
Wcltauffassung, 38 
Wenzl, Aloys, 189 
Wliilchcad, A. N., 265, 266, 300 
Wine, new, into old bottles, 25, 40 
WittRcnstein, L., 26, 31-32, 48, 103. 
112, 133, 256 

Words, Neurath’s index of prohibited, 
35; of everyday life, 244 
tVorfd, mechanistic and organismic con- 
ceptions of, 3, 4; scientific and mech- 
anistic viorvs of, 6-7, 58; structure of. 


in Einstein’s theory, 20; Gentral Eu- 
ropean movement toward scientific 
conception of, 26; apparent and real, 
39, 77. 80. 81, 134, 160, 163, 181, 
103; transcendental, 104; logical 
structure of. 111; outer, 112; mathe- 
matical, 123; organismic conception 
of, 126; built according to pure math- 
ematics, 132-134; mystical concep- 
tions of, 133; idealistic picture of, 
136; of atomic physics, 157; real 
metaphysical, 161; material, neglect 
of, 161; pin'sical, 243 
World architect, 133 
World creator, l33, 136 
World conception, in ctliical-rcligious 
sense, 60. Sec also Scientific world 
conception 

World pictnie, iSi); of priniitirc man, 
SO; pu'seientific ihcoiies as philo- 
sophic, 160 

World system. See Copernican system, 
Ptolemaic system 

World view, idealistic, 158, 186; de- 
cline of rational, 287 
World War 1, 18 


324